This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Many-body effects, orbital mixing and cyclotron resonance in bilayer graphene

K. Shizuya Yukawa Institute for Theoretical Physics
Kyoto University, Kyoto 606-8502, Japan
Abstract

In a magnetic field bilayer graphene supports, at the lowest Landau level, eight characteristic zero-energy levels with an extra twofold degeneracy in Landau orbitals n={0,1}n=\{0,1\}. They, under general one-body and many-body interactions, evolve into pseudo-zero-mode (PZM) levels. A close look is made into the detailed structure and characteristics of such PZM levels and cyclotron resonance they host, with full account taken of spin splitting, interlayer bias, weak electron-hole asymmetry and Coulomb interactions. It is pointed out that the PZM levels generally undergo orbital level mixing (in one valley or both valleys) as they are gradually filled with electrons and that an observation of interband cyclotron resonance over a finite range of filling provides a direct and sensitive probe for exploring many-body effects and orbital mixing in bilayer graphene.

I Introduction

Graphene supports as charge carriers massless Dirac fermions which display unique and fascinating electronic properties. Recently, increasing attention is directed to bilayers NMMKF ; MF and fewlayers of graphene, where the physics and applications of graphene become far richer, with, e.g., a tunable band gap MF ; OBSHR in bilayer graphene.

In a magnetic field, graphene leads to a ````relativistic” infinite tower of electron and hole Landau levels, with the lowest Landau level (LLL) consisting of a number of zero-energy levels. The presence of such zero-mode levels has a topological origin in the index of (the leading part of) the Dirac operators, or in the chiral anomaly NS . Monolayer graphene carries 2valley×2spin=42_{\rm valley}\times 2_{\rm spin}=4 zero-mode levels. Bilayer graphene supports an octet of such levels, with an extra twofold degeneracy in Landau orbitals n=0n=0 and n=1n=1. In real samples, topological zero-energy levels evolve, by an interplay of general spin and valley splittings and Coulomb interactions, into a variety of pseudo-zero-mode (PZM) levels, or broken-symmetry quantum Hall states, as discussed theoretically BCNM ; KSpzm ; BCLM ; CLBM ; CLPBM ; NL ; MAF . Meanwhile it was noted that orbital degeneracy is also lifted by Coulomb interactions alone KS_Ls : Quantum fluctuations of the valence band split, like the Lamb shift in the hydrogen atom, the n=0n=0 and n=1n=1 levels appreciably. The orbital degeneracy and its lifting by the orbital Lamb shift are new features specific to the LLL in fewlayer graphene. Experimentally, partial or full resolution of the eightfold degeneracy of the LLL in bilayer graphene has been observed in transport FMY ; ZCZJ ; WAFM ; MEM ; VJB and capacitance MFW ; KFLH ; HLZW measurements, but the detailed structure of the LLL remains yet to be clarified.

Graphene supports cyclotron resonance (CR) in a variety of channels, both intraband and interband AbF . Interband resonance is specific to Dirac electrons, with a number of competing resonance channels simultaneously activated over a certain range of the total Landau-level filling factor ν\nu. CR provides a useful means to look into the many-body problem in graphene. In experiment CR has been explored for monolayer JHTWS ; DCNN ; HCJL and bilayer HJTS ; OFBB graphene. Many-body effects on CR IWF ; BM ; KS_CR ; RFG ; KS_CR_eh ; ST ; KS_CRnewGR ; SL were first detected rather indirectly by a comparison of some leading intraband and interband resonances JHTWS . Recently, Russell et al. RZTW have reported a direct signal of many-body effects on CR in high-mobility hBN-encapsulated monolayer graphene; see also ref. prepH . They observed significant variations of resonance energies over a finite range of filling factor ν\nu under a fixed magnetic field BB; the point lies in extracting genuine Coulombic effects as a function of ν\nu while controlling the running of the velocity factor with BB. The data show a profile nearly symmetric in ν\nu and suggest a small band gap HuntYYY due to weak coupling to the hBN substrate.

No such ν\nu-dependent study of interband CR is yet available for bilayer graphene. The purpose of this paper is to examine, in anticipation of a future experiment, the detailed characteristics of the LLL in bilayer graphene, with full account taken of spin splitting, interlayer bias, weak electron-hole asymmetry and Coulomb interactions, and to show how they are detected by a ν\nu-dependent and fixed-BB survey of interband channels of CR. Particular attention is paid to electron-hole (ee-hh) conjugation and valley-interchange operations which relate the electron and hole bands within a valley or between the two valleys. We clarify how they govern the Landau-level and CR spectra.

In Sec. II we review some basic features of the effective theory of bilayer graphene in a magnetic field. In Sec. III we consider Coulombic corrections to Landau levels and CR. Such corrections inevitably contain ultraviolet divergences and, in Sec. IV, we carry out renormalization to extract observable many-body contributions. In Secs. V and VI, we examine how the PZM level spectra evolve with filling, note the general presence of orbital mixing and discuss how the associated many-body effects are revealed by interband CR. Section VII is devoted to a summary and discussion.

II bilayer graphene

The electrons in bilayer graphene are described by four-component spinor fields on the four inequivalent sites (A,B)(A,B) and (A,B)(A^{\prime},B^{\prime}) in the Bernal-stacked bottom and top layers of honeycomb lattices of carbon atoms. Their low-energy features are governed by the two Fermi points KK and KK^{\prime} in the Brillouin zone. The intralayer coupling γ0γAB3\gamma_{0}\equiv\gamma_{AB}\sim 3  eV is related to the Fermi velocity v=(3/2)aLγ0/106v=(\sqrt{3}/2)\,a_{\rm L}\gamma_{0}/\hbar\sim 10^{6} m/s (aL=0.246a_{\rm L}=0.246nm) in monolayer graphene. Interlayer hopping via the dimer coupling γ1γAB0.4\gamma_{1}\equiv\gamma_{A^{\prime}B}\sim 0.4 eV makes the spectra quasi-parabolic MF in the low-energy branches |ϵ|<γ1|\epsilon|<\gamma_{1}.

The bilayer Hamiltonian with γ0\gamma_{0} and γ1\gamma_{1} leads to ee-hh symmetric energy spectra. Infrared spectroscopy, however, detected weak asymmetry ZLBF ; LHJ_asym due to the (A,B)(A,B)-sublattice energy difference Δ18\Delta\sim 18meV and the nonleading interlayer coupling γ4γAA0.04γ0\gamma_{4}\equiv\gamma_{AA^{\prime}}\sim 0.04\gamma_{0}.

The effective Hamiltonian with such intra- and inter-layer couplings is written as MF ; NCGP

H\displaystyle H =\displaystyle= d2𝐱[(ΨK)KΨK+(ΨK)KΨK],\displaystyle\!\int\!d^{2}{\bf x}\,\Big{[}(\Psi^{K})^{{\dagger}}\,{\cal H}_{K}\Psi^{K}+(\Psi^{K^{\prime}})^{{\dagger}}\,{\cal H}_{K^{\prime}}\,\Psi^{K^{\prime}}\Big{]},
K\displaystyle{\cal H}_{K} =\displaystyle= (12u0v4pvp012uvpv4pv4pvpΔ12uγ1vpv4pγ1Δ+12u),\displaystyle\!\!\left(\begin{array}[]{cccc}{1\over{2}}u&0&-v_{4}\,p^{{\dagger}}&v\,p^{{\dagger}}\\ 0&-{1\over{2}}u&v\,p&-v_{4}\,p\\ -v_{4}\,p&v\,p^{{\dagger}}&\Delta-{1\over{2}}u&\gamma_{1}\\ v\,p&-v_{4}\,p^{{\dagger}}&\gamma_{1}&\Delta+{1\over{2}}u\\ \end{array}\right), (5)

with p=px+ipyp=p_{x}+i\,p_{y}, p=pxipyp^{{\dagger}}=p_{x}-i\,p_{y} and v4(γ4/γ0)vv_{4}\equiv(\gamma_{4}/\gamma_{0})\,v. Here ΨK=(ψB,ψA,ψB,ψA)t\Psi^{K}=(\psi_{B^{\prime}},\psi_{A},\psi_{B},\psi_{A^{\prime}})^{\rm t} denotes the electron field in valley KK, with AA and BB referring to the associated sublattices; uu stands for an interlayer bias, which opens a tunable valley gap OBSHR ; MF . We ignore the effect of trigonal warping γ3γAB0.1\propto\gamma_{3}\equiv\gamma_{AB^{\prime}}\sim 0.1 eV which, in a strong magnetic field, causes only a negligibly small level shift. K{\cal H}_{K} is diagonal in the (suppressed) electron spin.

As a typical set of band parameters we adopt those by recent theoretical calculations JM ; fn ,

v\displaystyle v \displaystyle\approx 0.845×106m/s,γ1361meV,\displaystyle 0.845\times 10^{6}\,{\rm m/s},\ \gamma_{1}\approx 361\,{\rm meV},
v4/v\displaystyle v_{4}/v \displaystyle\equiv γ4/γ00.053,15meV,\displaystyle\gamma_{4}/\gamma_{0}\approx 0.053,\ \triangle\approx 15\,{\rm meV}, (6)

which are also used in an experimental analysis HLZW .

Let us note that K{\cal H}_{K} is unitarily equivalent to K-{\cal H}_{K} with the signs of (u,Δ,v4)(u,\Delta,v_{4}) reversed,

SKS=K|u,Δ,v4,S^{{\dagger}}{\cal H}_{K}\,S=-{\cal H}_{K}|_{-u,-\Delta,-v_{4}}, (7)

with S=diag(1,1,1,1)S={\rm diag}(-1,1,-1,1); 𝒪|u,Δ,v4{\cal O}|_{-u,-\Delta,-v_{4}} signifies setting (u,Δ,v4)(u,Δ,v4)(u,\Delta,v_{4})\rightarrow(-u,-\Delta,-v_{4}) in 𝒪{\cal O}. The electron and hole bands are therefore intimately related within a valley.

The Hamiltonian K{\cal H}_{K^{\prime}} in another valley is given by K{\cal H}_{K} with (v,v4,u)(v,v4,u)(v,v_{4},u)\rightarrow(-v,-v_{4},-u), and acts on a spinor of the form ΨK=(ψA,ψB,ψA,ψB)t\Psi^{K^{\prime}}=(\psi_{A},\psi_{B^{\prime}},\psi_{A^{\prime}},\psi_{B})^{\rm t}. Actually, K{\cal H}_{K^{\prime}} is unitarily equivalent to K{\cal H}_{K} with the sign of uu reversed,

K=𝒰K|u𝒰,{\cal H}_{K^{\prime}}={\cal U}^{{\dagger}}{\cal H}_{K}|_{-u}\,{\cal U}, (8)

with 𝒰=diag(1,1,1,1){\cal U}={\rm diag}(1,1,-1,-1). In what follows we adopt K|u{\cal H}_{K}|_{-u} for K{\cal H}_{K^{\prime}} and simply pass to valley KK^{\prime} by reversing the sign of bias uu in valley-KK expressions. We also suppose, without loss of generality, that u0u\geq 0 in valley KK; thus u0u\leq 0 refers to valley KK^{\prime}.

The unitary equivalence

K𝒰K|uSK|Δ,v4{\cal H}_{K^{\prime}}\stackrel{{\scriptstyle\cal U}}{{\sim}}{\cal H}_{K}|_{-u}\stackrel{{\scriptstyle S}}{{\sim}}-{\cal H}_{K}|_{-\Delta,-v_{4}} (9)

implies that the ee-hh conjugation symmetry is kept exact only for Δ=v4=0\Delta=v_{4}=0 although, if u0u\not=0, it is apparently broken within each valley. This symmetry analysis also holds in the presence of Coulomb interactions.

Let us place bilayer graphene in a strong uniform magnetic field Bz=B>0B_{z}=B>0 normal to the sample plane; we set, in K{\cal H}_{K}, pΠ=p+eAp\rightarrow\Pi=p+eA with A=Ax+iAy=ByA=A_{x}+iA_{y}=-By, and rescale Π=(2/)Z\Pi=-(\sqrt{2}/\ell)\,Z^{{\dagger}} so that [Z,Z]=1[Z,Z^{{\dagger}}]=1, where =1/eB\ell=1/\sqrt{eB} denotes the magnetic length.

The eigenmodes of K{\cal H}_{K} are labeled by integers N|n|=(0,1,2,)N\equiv|n|=(0,1,2,\cdots) and plane waves with momentum pxp_{x}, and have the structure

Ψn=(|N2bn,|Ncn,|N1bn,|N1cn)t,\Psi^{n}=\Big{(}|N\!-\!2\rangle\,b^{\prime}_{n},|N\rangle\,c_{n},|N\!-\!1\rangle\,b_{n},|N\!-\!1\rangle\,c^{\prime}_{n}\Big{)}^{\rm t}, (10)

where only the orbital modes are shown using the standard harmonic-oscillator basis {|N}\{|N\rangle\}, with |N=0|N\rangle=0 for N<0N<0. The coefficients 𝐛n=(bn,cn,bn,cn)t{\bf b}_{n}=(b^{\prime}_{n},c_{n},b_{n},c^{\prime}_{n})^{\rm t} for each NN are given by the normalized eigenvectors of the reduced Hamiltonian

^N=ωc(μ0rN1N10μNrNrN1NdμgN1rNgd+μ),\hat{\cal H}_{N}=\omega_{c}\left(\begin{array}[]{cccc}\mu&0&r\,\sqrt{N\!-\!1}&-\sqrt{N\!-\!1}\\ 0&-\mu&-\sqrt{N}&r\,\sqrt{N}\\ r\,\sqrt{N\!-\!1}&-\sqrt{N}&d-\mu&g\\ -\sqrt{N\!-\!1}&r\,\sqrt{N}&g&d+\mu\\ \end{array}\right), (11)

where

ωc2v/36.3×v[106m/s]B[T]meV,\omega_{c}\equiv\sqrt{2}\,v/\ell\approx 36.3\times v[10^{6}{\rm m/s}]\,\sqrt{B[{\rm T}]}\ {\rm meV}, (12)

is the cyclotron energy for monolayer graphene; gγ1/ωcg\equiv\gamma_{1}/\omega_{c}, μ12u/ωc\mu\equiv{1\over{2}}\,u/\omega_{c}, dΔ/ωcd\equiv\Delta/\omega_{c} and rγ4/γ0r\equiv\gamma_{4}/\gamma_{0}. Numerically, for the set of parameters in Eq. (6) at B=20B=20 T,

ωc137meV,(g,d,r)(2.63,0.109,0.053),\omega_{c}\approx 137\,{\rm meV},\ (g,d,r)\approx(2.63,0.109,0.053), (13)

and μ0.073\mu\approx 0.073 for u=20u=20meV. In what follows, we employ this set (13) of parameters for numerical estimates at B=20B=20 T.

The unitary equivalence in Eq. (9) reveals how the spectra ϵn\epsilon_{n} and the associated eigenmodes 𝐛n{\bf b}_{n} change via ee-hh conjugation. Within each valley they read

ϵn=ϵn|u,Δ,v4,\displaystyle\epsilon_{-n}=-\epsilon_{n}|_{-u,-\Delta,-v_{4}},
(bn,cn,bn,cn)=(bn,cn,bn,cn)|u,Δ,v4,\displaystyle(b^{\prime}_{-n},c_{-n},b_{-n},c^{\prime}_{-n})=(-b^{\prime}_{n},c_{n},-b_{n},c^{\prime}_{n})|_{-u,-\Delta,-v_{4}},\ \ \ \ (14)

while, between the two valleys, they are related as

(ϵn;bn,cn,bn,cn)|K=(ϵn;bn,cn,bn,cn)|uK\displaystyle(\epsilon_{n};b^{\prime}_{n},c_{n},b_{n},c^{\prime}_{n})|^{K^{\prime}}=(\epsilon_{n};b^{\prime}_{n},c_{n},b_{n},c^{\prime}_{n})|^{K}_{-u}
=(ϵn;bn,cn,bn,cn)|Δ,v4K.\displaystyle=(-\epsilon_{-n};b^{\prime}_{-n},c_{-n},-b_{-n},c^{\prime}_{-n})|^{K}_{-\Delta,-v_{4}}. (15)

Of our particular concern are the n=0n=0 and n=1n=1 modes. For n=0n=0, ^N=0\hat{\cal H}_{N=0} has an obvious eigenvalue and eigenvector

ϵ0=u/2=ωcμ,𝐛0=(0,1,0,0)t.\epsilon_{0}=-u/2=-\omega_{c}\mu,\ \ {\bf b}_{0}=(0,1,0,0)^{\rm t}. (16)

For N=1N=1, ^N\hat{\cal H}_{N} has three solutions (ϵ1,ϵ1,ϵ1)(\epsilon_{-1^{\prime}},\epsilon_{1},\epsilon_{1^{\prime}}), with ϵ±1±γ1\epsilon_{\pm 1^{\prime}}\sim\pm\gamma_{1} belonging to the higher branches.

The n=1n=1 mode (with ϵ1\epsilon_{1}) and the n=0n=0 mode have zero energy for (u,Δ,r)=0(u,\Delta,r)=0 and deviate from zero energy as (u,Δ,r)(u,\Delta,r) develop. These pseudo-zero modes (PZM) form the LLL in bilayer graphene. The eigenenergy and eigenvector of the n=1n=1 mode are written as

ϵ1\displaystyle\epsilon_{1} =\displaystyle= ωc(μ+κ),\displaystyle\omega_{c}(-\mu+\kappa),
𝐛1\displaystyle{\bf b}_{1} =\displaystyle= c1(0,1,(gκr)/g~,(1+κdd2)/g~),\displaystyle c_{1}\Big{(}0,1,-(g\kappa-r)/\tilde{g},(1+\kappa d-d^{2})/\tilde{g}\Big{)},\ \ \
g~\displaystyle\tilde{g} =\displaystyle= g+(dκ)r.\displaystyle g+(d-\kappa)\,r. (17)

In what follows, we denote valley and ee-hh symmetry breaking collectively as X=(u,Δ,v4)X=(u,\Delta,v_{4}) or X=(μ,d,r)X=(\mu,d,r). As seen from the secular equation, the correction κκ(g;μ,d,r)\kappa\equiv\kappa(g;\mu,d,r) in ϵ1\epsilon_{1} is odd in XX, and hence the normalization factor c1c_{1} is even in XX,

κ\displaystyle\kappa \displaystyle\approx 2μ+d+2rgg2+1+O(X3),\displaystyle{2\mu+d+2r\,g\over{g^{2}+1}}+O(X^{3}),
c1\displaystyle c_{1} \displaystyle\approx 1/1+(1/g2)+O(X2).\displaystyle 1/\sqrt{1+(1/g^{2})}+O(X^{2}). (18)

Numerically, at B=20B=20 T, κ0.049+0.25μ\kappa\approx 0.049+0.25\,\mu while c10.9340.007μ0.02μ2c_{1}\approx 0.934-0.007\,\mu-0.02\,\mu^{2} scarcely depends on μ\mu [for small μO(0.1)\mu\sim O(0.1)].

Interlayer bias uμu\propto\mu, if nonzero, breaks valley symmetry and shifts the PZM levels n={0,1}n=\{0,1\} oppositely (u/2\propto\mp u/2) in the two valleys. Accordingly, (ϵ0,ϵ1)(\epsilon_{0},\epsilon_{1}) are nearly degenerate in each valley and are ordered so that ϵ0Kϵ1K<0<ϵ1Kϵ0K\epsilon_{0}^{K}\lesssim\epsilon_{1}^{K}<0<\epsilon_{1}^{K^{\prime}}\lesssim\epsilon_{0}^{K^{\prime}} for μ>0\mu>0.

For N2N\geq 2, ^N\hat{\cal H}_{N} has rank 4 and we denote the four branches of Landau-level spectra as ϵn<ϵn<0<ϵn<ϵn\epsilon_{-n^{\prime}}<\epsilon_{-n}<0<\epsilon_{n}<\epsilon_{n^{\prime}} (with |ϵ±n|γ1|\epsilon_{\pm n^{\prime}}|\gtrsim\gamma_{1}) so that the index (n,n)(n,n^{\prime}) reflects the sign of ϵn\epsilon_{n}. Let us denote ϵn\epsilon_{n} in units of ωc\omega_{c},

ϵn=ωcenandϵn=ωcen.\epsilon_{n}=\omega_{c}\,e_{n}\ {\rm and}\ \epsilon_{n^{\prime}}=\omega_{c}\,e_{n^{\prime}}. (19)

For X=(μ,d,r)0X=(\mu,d,r)\rightarrow 0, the (dimensionless) spectra enen(g;μ,d,r)e_{n}\equiv e_{n}(g;\mu,d,r) of the lower branches take the form

en(0)=sn2|n|(|n|1)g2+an+DnO(1/g),e_{n}^{(0)}=s_{n}\sqrt{{2|n|(|n|-1)\over{g^{2}+a_{n}+\sqrt{D_{n}}}}}\sim O(1/g), (20)

where en(0)en|X=0e_{n}^{(0)}\equiv e_{n}|_{X=0}, Dn=g4+2ang2+1D_{n}=g^{4}+2a_{n}g^{2}+1 and an=2|n|1a_{n}=2\,|n|-1; snsgn[n]=±1s_{n}\equiv{\rm sgn}[n]=\pm 1 is the sign function. The full spectra, to first order in breaking XX, are written as

enen(0)+12(1g2Dn)d+μ+anrgDn.e_{n}\approx e_{n}^{(0)}+{1\over{2}}(1-{g^{2}\over{\sqrt{D_{n}}}})\,d+{\mu+a_{n}\,r\,g\over{\sqrt{D_{n}}}}. (21)

Actually, ee-hh breaking (d,r)(d,r), listed in Eq. (13), has a sizable effect and shifts all the spectra, except for ϵ0\epsilon_{0}, upwards appreciably [e.g., roughly by 10 % for n=(2,3)n=(2,3)], as seen from Fig. 1, which, for later convenience, is presented in Sec. V. Unlike e0e1O(μ)e_{0}\sim e_{1}\sim O(\mu), ene_{n} depend on bias uu only weakly O(μ/g2)\sim O(\mu/g^{2}). The spectra ϵn\epsilon_{n^{\prime}} of the higher branches n=±1,±2,n^{\prime}=\pm 1^{\prime},\pm 2^{\prime},\dots show similar but partially different dependence on breaking XX.

Each |n|2|n|\geq 2 is associated with a pair of electron and hole modes ±n\pm n (and ±n\pm n^{\prime}). In contrast, the pseudo-zero modes n=0n=0 and n=1n=1 stand alone (per spin and valley) and are, in this sense, ee-hh selfconjugate, with ±nn\pm n\rightarrow n in Eqs. (14) and (15). Thus ϵ0=ϵ0|X=ωcμ\epsilon_{0}=-\epsilon_{0}|_{-X}=-\omega_{c}\mu and ϵ1=ϵ1|X\epsilon_{1}=-\epsilon_{1}|_{-X}; that is, ϵ0=ϵ1=0\epsilon_{0}=\epsilon_{1}=0 for X0X\rightarrow 0 and ϵ1\epsilon_{1} consists of odd powers of breaking XX.

The Landau-level structure is made explicit by passing to the |n,y0|n,y_{0}\rangle basis (with y02pxy_{0}\equiv\ell^{2}p_{x}) via the expansion Ψαa(𝐱)=n,y0𝐱|n,y0ψαn;a(y0)\Psi^{a}_{\alpha}({\bf x})=\sum_{n,y_{0}}\langle{\bf x}|n,y_{0}\rangle\,\psi^{n;a}_{\alpha}(y_{0}), where nn refers to the Landau level index, α(,)\alpha\in(\uparrow,\downarrow) to the spin and a(K,K)a\in(K,K^{\prime}) to the valley. It is tacitly understood that the sum is taken over the higher branches n=(±1,±2,)n^{\prime}=(\pm 1^{\prime},\pm 2^{\prime},\dots) as well. The one-body Hamiltonian is written as

H=𝑑y0m,na,αψαm;a(y0)ϵna;αψαn;a(y0).H=\int dy_{0}\sum_{m,n}\sum_{a,\alpha}\psi_{\alpha}^{m;a{\dagger}}(y_{0})\,\epsilon_{n}^{a;\alpha}\,\psi_{\alpha}^{n;a}(y_{0}). (22)

The charge density ρ𝐩=d2𝐱ei𝐩𝐱ρ\rho_{-{\bf p}}=\int d^{2}{\bf x}\,e^{i{\bf p\cdot x}}\,\rho with ρ=(ΨK)ΨK+(ΨK)ΨK\rho=(\Psi^{K})^{{\dagger}}\Psi^{K}+(\Psi^{K^{\prime}})^{{\dagger}}\Psi^{K^{\prime}} is thereby written as KS_CR

ρ𝐩\displaystyle\rho_{-{\bf p}} =\displaystyle= γ𝐩m,n=a,αg𝐩mn;aRαα;𝐩mn;aa,\displaystyle\gamma_{\bf p}\sum_{m,n=-\infty}^{\infty}\sum_{a,\alpha}g^{mn;a}_{\bf p}\,R^{mn;aa}_{\alpha\alpha;\bf-p},
Rαβ;𝐩mn;ab\displaystyle R^{mn;ab}_{\alpha\beta;{\bf-p}} \displaystyle\equiv 𝑑y0ψαm;a(y0)ei𝐩𝐫ψβn;b(y0),\displaystyle\int dy_{0}\,{\psi^{m;a}_{\alpha}}^{{\dagger}}(y_{0})\,e^{i{\bf p\cdot r}}\,\psi^{n;b}_{\beta}(y_{0}), (23)

where γ𝐩=e2𝐩2/4\gamma_{\bf p}=e^{-\ell^{2}{\bf p}^{2}/4}; 𝐫=(i2/y0,y0){\bf r}=(i\ell^{2}\partial/\partial y_{0},y_{0}) stands for the center coordinate with uncertainty [rx,ry]=i2[r_{x},r_{y}]=i\ell^{2}. The charge operators Rαα;𝐩mn;aaR^{mn;aa}_{\alpha\alpha;\bf p} obey the WW_{\infty} algebra GMP .

The coefficient matrix g𝐩mn;ag𝐩mn|ag^{mn;a}_{\bf p}\equiv g^{mn}_{\bf p}|^{a} in valley a(K,K)a\in(K,K^{\prime}) is constructed from the eigenvectors 𝐛n|a{\bf b}_{n}|^{a},

g𝐩mn\displaystyle g^{mn}_{\bf p} =\displaystyle= cmcnf𝐩|m|,|n|+bmbnf𝐩|m|2,|n|2\displaystyle c_{m}\,c_{n}\,f_{\bf p}^{|m|,|n|}+b^{\prime}_{m}\,b^{\prime}_{n}\,f_{\bf p}^{|m|-2,|n|-2} (24)
+(bmbn+cmcn)f𝐩|m|1,|n|1,\displaystyle+(b_{m}\,b_{n}+c^{\prime}_{m}\,c^{\prime}_{n})\,f_{\bf p}^{|m|-1,|n|-1},

where

f𝐩mn=n!m!(p2)mnLn(mn)(122𝐩2)f^{mn}_{\bf p}=\sqrt{{n!\over{m!}}}\,\Big{(}{-\ell p^{{\dagger}}\over{\sqrt{2}}}\Big{)}^{m-n}\,L^{(m-n)}_{n}\Big{(}{1\over{2}}\ell^{2}{\bf p}^{2}\Big{)} (25)

for mn0m\geq n\geq 0, and f𝐩nm=(f𝐩mn)f^{nm}_{\bf p}=(f^{mn}_{\bf-p})^{{\dagger}}; f𝐩mn=0f^{mn}_{\bf p}=0 for m<0m<0 or n<0n<0; p=px+ipyp=p_{x}\!+i\,p_{y}. In view of Eqs. (14) and  (15), g𝐩mng^{mn}_{\bf p} have the following property under ee-hh conjugation,

g𝐩m,n;a=g𝐩m,n;a|μ,d,r,\displaystyle g^{-m,-n;a}_{\bf p}=g^{m,n;a}_{\bf p}|_{-\mu,-d,-r},
g𝐩mn|K=g𝐩m,n|μK=g𝐩m,n|d,rK.\displaystyle g^{mn}_{\bf p}|^{K^{\prime}}=g^{m,n}_{\bf p}|^{K}_{-\mu}=g^{-m,-n}_{\bf p}|^{K}_{-d,-r}. (26)

For n,m(0,1)n,m\in(0,1), g𝐩m,n;a=g𝐩m,n;a|Xg^{m,n;a}_{\bf p}=g^{m,n;a}_{\bf p}|_{-X} are even functions of breaking XX. They actually take simple form

g𝐩00\displaystyle g^{00}_{\bf p} =\displaystyle= 1,g𝐩11=1c12122𝐩2,\displaystyle 1,\ \ g^{11}_{\bf p}=1-c_{1}^{2}\,\textstyle{1\over{2}}\ell^{2}{\bf p}^{2},
g𝐩01\displaystyle g^{01}_{\bf p} =\displaystyle= c1p/2,g𝐩10=c1p/2,\displaystyle c_{1}\ell\,p/\sqrt{2},\ \ g^{10}_{\bf p}=-c_{1}\ell\,p^{{\dagger}}/\sqrt{2}, (27)

in each valley, with c1|K=c1|μKc_{1}|^{K^{\prime}}=c_{1}|^{K}_{-\mu}.

The Coulomb interaction V=12𝐩v𝐩:ρ𝐩ρ𝐩:V={1\over{2}}\sum_{\bf p}v_{\bf p}\,{:\!\rho_{\bf-p}\,\rho_{\bf p}\!:} is written as

V=12𝐩v𝐩γ𝐩2g𝐩jk;ag𝐩mn;b:Rαα;𝐩jk;aaRββ;𝐩mn;bb:,V={1\over{2}}\sum_{\bf p}v_{\bf p}\,\gamma_{\bf p}^{2}\,g^{jk;a}_{\bf p}\,g^{mn;b}_{\bf-p}:\!R^{jk;aa}_{\alpha\alpha;{\bf-p}}\,R^{mn;bb}_{\beta\beta;{\bf p}}\!:, (28)

with the potential v𝐩=2παe/(ϵb|𝐩|)v_{\bf p}=2\pi\alpha_{e}/(\epsilon_{\rm b}|{\bf p}|), αee2/(4πϵ0)\alpha_{e}\equiv e^{2}/(4\pi\epsilon_{0}) and the substrate dielectric constant ϵb\epsilon_{\rm b}; 𝐩d2𝐩/(2π)2\sum_{\bf p}\equiv\int d^{2}{\bf p}/(2\pi)^{2}; :::\ : denotes normal ordering. For simplicity, we ignore a small effect of interlayer separation. Here and from now on, we suppress summations over levels nn, valleys aa and spins α\alpha, with the convention that the sum is taken over repeated indices. The one-body Hamiltonian HH is thereby written as

H=ϵna;αRαα;𝐩=𝟎nn;aaμZ(T3)αβRαβ;𝐩=𝟎nn;aa.H=\epsilon_{n}^{a;\alpha}\,R^{nn;aa}_{\alpha\alpha;{\bf p=0}}-\mu_{\rm Z}\,(T_{3})_{\alpha\beta}R^{nn;aa}_{\alpha\beta;{\bf p=0}}. (29)

Here the Zeeman term μZgμBB\mu_{\rm Z}\equiv g^{*}\mu_{\rm B}B is introduced via the spin matrix T3=σ3/2T_{3}=\sigma_{3}/2.

III Coulombic corrections

In this section we study Coulombic contributions to the Landau-level and associated CR spectra. In graphene, unlike conventional quantum Hall systems, the electrons and holes are always subject to quantum fluctuations of the infinitely-deep filled valence band (or the Dirac sea), which, being strong, lead to ultraviolet divergences. One first has to handle them properly.

The Coulomb direct interaction leads to a divergent self-energy v𝐩𝟎\propto v_{\bf p\rightarrow 0}, which, as usual, is removed when a neutralizing background is taken into account. The exchange interaction, on the other hand, gives rise to corrections to level spectra ϵna;α\epsilon_{n}^{a;\alpha} of the form

Δϵna;α=𝐩v𝐩γ𝐩2mνma;α|g𝐩nm;a|2.\Delta\epsilon_{n}^{a;\alpha}=-\sum_{\bf p}v_{\bf p}\gamma_{\bf p}^{2}\,\sum_{m}\nu_{m}^{a;\alpha}\,|g^{nm;a}_{\bf p}|^{2}. (30)

Here 0νna;α10\leq\nu_{n}^{a;\alpha}\leq 1 stands for the filling fraction of the (n,a,α)(n,a,\alpha) level. The exchange interaction preserves the spin and valley (α,a)(\alpha,a). Accordingly, from now on, we suppress them and mainly refer to valley KK.

The self-energies Δϵn\Delta\epsilon_{n} involve a sum over infinitely many filled levels mm in the valence band. Their structure is clarified if one notes the completeness relation

k=|g𝐩nk|2=1/γ𝐩2=e122𝐩2.\sum_{k=-\infty}^{\infty}|g^{nk}_{\bf p}|^{2}=1/\gamma_{\bf p}^{2}=e^{{1\over{2}}\ell^{2}{\bf p}^{2}}. (31)

Actually, this follows from the fact KS_LWGS that G𝐩nkγ𝐩g𝐩nkG^{nk}_{\bf p}\equiv\gamma_{\bf p}\,g^{nk}_{\bf p} are (WW_{\infty}) unitary matrices that obey the composition law G𝐩G𝐪=ei122𝐩×𝐪G𝐩+𝐪G_{\bf p}\,G_{\bf q}=e^{i{1\over{2}}\ell^{2}{\bf p}\times{\bf q}}\,G_{\bf p+q} with 𝐩×𝐪=pxqypyqx{\bf p}\times{\bf q}=p_{x}q_{y}-p_{y}q_{x}. The half-infinite sum in Δϵn\Delta\epsilon_{n} is then cast in the form

γ𝐩2k2|g𝐩nk|2\displaystyle\gamma_{\bf p}^{2}\!\sum_{k\leq-2}|g^{nk}_{\bf p}|^{2}\! =\displaystyle= 1212Fn(z)12γ𝐩2(|g𝐩n0|2+|g𝐩n1|2),\displaystyle\!{\textstyle{1\over{2}}}-{\textstyle{1\over{2}}}\,F_{n}(z)-{\textstyle{1\over{2}}}\gamma_{\bf p}^{2}\,(|g^{n0}_{\bf p}|^{2}+|g^{n1}_{\bf p}|^{2}),
Fn(z)\displaystyle F_{n}(z) \displaystyle\equiv γ𝐩2k=2{|g𝐩nk|2|g𝐩n,k|2},\displaystyle\gamma_{\bf p}^{2}\sum_{k=2}^{\infty}\{|g^{nk}_{\bf p}|^{2}-|g^{n,-k}_{\bf p}|^{2}\}, (32)

where z=122𝐩2z={1\over{2}}\ell^{2}{\bf p}^{2}. Here again it is tacitly understood that the sum is taken over k1k^{\prime}\leq-1 or k1k^{\prime}\geq 1 as well.

The self-energies Δϵn\Delta\epsilon_{n} are thereby rewritten as

Δϵn=𝐩v𝐩[12+12Fn(z)kν[k]γ𝐩2|g𝐩nk|2].\Delta\epsilon_{n}=\sum_{\bf p}v_{\bf p}\Big{[}\!-{\textstyle{1\over{2}}}+{\textstyle{1\over{2}}}\,F_{n}(z)-\sum_{k}\nu[k]\,\gamma_{\bf p}^{2}|g^{nk}_{\bf p}|^{2}\Big{]}. (33)

Here the last term with the ````electron-hole” filling factor,

ν[k]\displaystyle\nu[k] =\displaystyle= νkθ(k2)(1νk)θ(k2)\displaystyle\nu_{k}\,\theta_{(k\geq 2)}-(1-\nu_{k})\theta_{(k\leq-2)} (34)
+(ν112)δk,1+(ν012)δk,0,\displaystyle+(\nu_{1}-{\textstyle{1\over{2}}})\,\delta_{k,1}+(\nu_{0}-{\textstyle{1\over{2}}})\,\delta_{k,0},

where θ(k2)=1\theta_{(k\geq 2)}=1 for k2k\geq 2 and θ(k2)=0\theta_{(k\geq 2)}=0 otherwise, etc., refers to a finite number of filled electron or hole levels around the PZM sector n={0,1}n=\{0,1\}.

Now Fn(z)Fn(z;g,μ,d,r)F_{n}(z)\equiv F_{n}(z;g,\mu,d,r) involve a sum over infinitely many levels. Fn(z)1/zF_{n}(z)\propto 1/\sqrt{z} for zz\rightarrow\infty and yield ultraviolet divergences upon integration over 𝐩{\bf p} with v𝐩v_{\bf p}. In view of Eq. (26), F±n(z)F_{\pm n}(z) are related in each valley or between the valleys,

Fn(z)\displaystyle F_{-n}(z) =\displaystyle= Fn(z)|μ,d,r,\displaystyle-F_{n}(z)|_{-\mu,-d,-r},
Fn(z)|K\displaystyle F_{n}(z)|^{K^{\prime}} =\displaystyle= Fn(z)|μK=Fn(z)|d,rK.\displaystyle F_{n}(z)|^{K}_{-\mu}=-F_{-n}(z)|^{K}_{-d,-r}. (35)

For n=(0,1)n=(0,1), Fn(z)=Fn(z)|XF_{n}(z)=-F_{n}(z)|_{-X}; F0(z)F_{0}(z) and F1(z)F_{1}(z) are odd in breaking X=(u,d,r)X=(u,d,r). The PZM sector thus has no divergence for X=0X=0, (F0(z),F1(z))|X00(F_{0}(z),F_{1}(z))|_{X\rightarrow 0}\rightarrow 0.

Let us eliminate from Δϵn\Delta\epsilon_{n} a constant, 𝐩v𝐩[12+12F0(z)|μ=0]\sum_{\bf p}v_{\bf p}\,[-{1\over{2}}+{1\over{2}}F_{0}(z)|_{\mu=0}] common to all levels, which is safely done by adjusting zero of energy. Via such regularization, self-energies Δϵn\Delta\epsilon_{n} are cast in the form

Δϵn=reg12𝐩v𝐩n(z)kν[k]𝐩v𝐩γ𝐩2|g𝐩nk|2.\Delta\epsilon_{n}\stackrel{{\scriptstyle\rm reg}}{{=}}{\textstyle{1\over{2}}}\sum_{\bf p}v_{\bf p}\,{\cal F}_{n}(z)-\sum_{k}\nu[k]\sum_{\bf p}v_{\bf p}\,\gamma_{\bf p}^{2}|g^{nk}_{\bf p}|^{2}. (36)

Note that

n(z)Fn(z)F0(z)|μ=0.{\cal F}_{n}(z)\equiv F_{n}(z)-F_{0}(z)|_{\mu=0}. (37)

inherit the conjugation property of Fn(z)F_{n}(z) in Eq. (35).

A key property of the electron-hole filling factor ν[k]\nu[k] defined in Eq. (34) is that it is odd under ee-hh conjugation: ν[k]\nu[k] changes sign upon interchanging the electron and hole levels (kk)(k\rightarrow-k) by replacing, in ν[k]\nu[-k], νk1νk\nu_{-k}\rightarrow 1-\nu_{k}. Noting this feature and rewriting Eq. (36) then allows us to relate, as done earlier KS_CRnewGR for monolayer graphene, Δϵ±n\Delta\epsilon_{\pm n} in a valley or between the valleys. In terms of the full spectra to O(V)O(V),

ϵ^n=ϵn+Δϵn,\hat{\epsilon}_{n}=\epsilon_{n}+\Delta\epsilon_{n}, (38)

the result is

ϵ^nK|Nf=m=ϵ^nK|μ,d,rNf=2m=ϵ^nK|d,rNf=2m.\hat{\epsilon}_{n}^{K}|^{N_{\rm f}=m}=-\hat{\epsilon}_{-n}^{K}|^{N_{\rm f}=2-m}_{-\mu,-d,-r}=-\hat{\epsilon}_{-n}^{K^{\prime}}|^{N_{\rm f}=2-m}_{-d,-r}. (39)

Here NfN_{\rm f} specifies filling of the relevant valley; note that Δϵn\Delta\epsilon_{n} depend on filling. We assign Nf=0N_{\rm f}=0 to the empty PZM sector (of a given spin and valley) with levels below it (n2n\leq-2) all filled, or with the ````uppermost filled level” nf=2n_{\rm f}=-2; accordingly, Nf=(1,0;3,4)N_{\rm f}=(-1,0;3,4), e.g., refer to nf=(3,2;2,3)n_{\rm f}=(-3,-2;2,3). Nf=1N_{\rm f}=1 refers to the PZM sector with one filled level and Nf=2N_{\rm f}=2 to the filled sector. For definiteness, in what follows, we focus on cases of integer filling, in which a distinct band gap is present, and take NfN_{\rm f} in Eq. (39) to be integers. Let us now imagine, e.g., valley KK with Nf=mN_{\rm f}=m. Interchanging electrons and holes (according to νk1νk\nu_{-k}\rightarrow 1-\nu_{k}) yields a configuration with Nf=2mN_{\rm f}=2-m, and Eq. (39) implies that the spectra of such ee-hh conjugate configurations are intimately related. A typical ee-hh conjugate pair are the empty PZM sector (Nf=0N_{\rm f}=0) and the filled one (Nf=2N_{\rm f}=2).

Some care is needed in handling the Nf=1N_{\rm f}=1 configuration, which is not uniquely fixed, e.g., Nf=1N_{\rm f}=1 with nf=0n_{\rm f}=0 or nf=1n_{\rm f}=1 or with any linear combination of {0,1}\{0,1\}. Note also that ee-hh conjugation reverses the level layout of the PZM sector, e.g., (0,1)(1,0)(0,1)\rightarrow(1,0). As a result, if, e.g., ϵ^0<ϵ^1\hat{\epsilon}_{0}<\hat{\epsilon}_{1} in a valley, Eq. (39) reads

(ϵ^0,ϵ^1)|Nf=1,nf=0=(ϵ^0,ϵ^1)|XNf=1,nf=1,(\hat{\epsilon}_{0},\hat{\epsilon}_{1})|^{N_{\rm f}=1,n_{\rm f}=0}=(-\hat{\epsilon}_{0},-\hat{\epsilon}_{1})|^{N_{\rm f}=1,n_{\rm f}=1}_{-X}, (40)

which means that the filled n=0n=0 level turns into the filled n=1n=1 level in the conjugate configuration. In Sec. V, we study the PZM sector over a continuous range 0Nf20\leq N_{\rm f}\leq 2 and encounter an interesting case of mixed {0,1}\{0,1\} levels at Nf=1N_{\rm f}=1.

Let us next study CR, namely, optical interlevel transitions at zero momentum transfer, with the selection rule AbF Δ|n|=±1\Delta|n|=\pm 1 for graphene, i.e., (i) intraband channels {nn1,(n1)n}\{n\leftarrow n-1,-(n-1)\leftarrow-n\} and (ii) interband channels Tn{n(n1),n1n}T_{n}\equiv\{n\leftarrow-(n-1),\ n-1\leftarrow-n\} for n=1,2,n=1,2,\cdots; T1={10}T_{1}=\{1\!\leftrightarrow\!0\}, T2={21, 12}T_{2}=\{2\leftarrow 1,\ 1\leftarrow-2\}, T3={32, 23}T_{3}=\{3\leftarrow-2,\ 2\leftarrow-3\}, etc. Interband CR is specific to Dirac electrons. Consider now CR from level jj to level nn for each (valley, spin)=(a,α)(a,\alpha) channel and denote the associated excitation energy as ϵexcnj=ϵnϵj+Δϵn,j\epsilon_{\rm exc}^{n\leftarrow j}=\epsilon_{n}-\epsilon_{j}+\Delta\epsilon^{n,j}; CR preserves the valley and spin so that (a,α)(a,\alpha) will be suppressed from now on. In the mean-field treatment KS_CR ; KH ; MOG , the corrections

Δϵn,j\displaystyle\Delta\epsilon^{n,j} =\displaystyle= ΔϵnΔϵj(νjνn)𝒜n,j,\displaystyle\Delta\epsilon_{n}-\Delta\epsilon_{j}-(\nu_{j}-\nu_{n})\,{\cal A}_{n,j},
𝒜n,j\displaystyle{\cal A}_{n,j} =\displaystyle= 𝐩v𝐩γ𝐩2g𝐩nng𝐩jj,\displaystyle\sum_{\bf p}v_{\bf p}\gamma_{\bf p}^{2}\,g^{nn}_{\bf-p}\,g^{jj}_{\bf p}, (41)

consist of the self-energy difference and the Coulombic attraction 𝒜n,j{\cal A}_{n,j} between the excited electron and created hole. The full CR spectra thus differ from the dressed level gaps by attraction energies 𝒜n,j{\cal A}_{n,j},

ϵexcnj=ϵ^nϵ^j(νjνn)𝒜n,j.\epsilon_{\rm exc}^{n\leftarrow j}=\hat{\epsilon}_{n}-\hat{\epsilon}_{j}-(\nu_{j}-\nu_{n})\,{\cal A}_{n,j}. (42)

It will be clear now that, via ee-hh conjugation, ϵexcnj\epsilon_{\rm exc}^{n\leftarrow-j} turns into ϵexcjn\epsilon_{\rm exc}^{j\leftarrow-n}. In particular, the interband CR channels Tn{n(n1),n1n}T_{n}\equiv\{n\leftarrow-(n-1),\ n-1\leftarrow-n\} are intimately related,

ϵexcn(n1)|K;Nf=m\displaystyle\epsilon_{\rm exc}^{n\leftarrow-(n-1)}|^{K;N_{\rm f}=m}\! =\displaystyle= ϵexcn1n|μ,d,rK;Nf=2m,\displaystyle\epsilon_{\rm exc}^{n-1\leftarrow-n}|^{K;N_{\rm f}=2-m}_{-\mu,-d,-r}, (43)
=\displaystyle= ϵexcn1n|d,rK;Nf=2m;\displaystyle\epsilon_{\rm exc}^{n-1\leftarrow-n}|^{K^{\prime};N_{\rm f}=2-m}_{-d,-r};

we suppose Nf1N_{\rm f}\not=1 here. Such an ````ee-hh conjugate” character of TnT_{n} was noticed earlier KS_CRnewGR for monolayer graphene, for which ee-hh conjugation is an exact symmetry. Here for bilayer graphene ee-hh conjugation, though not a symmetry, is an operation that tells us how the CR spectra deviate from the symmetric ones by breaking (μ,d,r)(\mu,d,r). Experimentally the conjugate channels of a given set TnT_{n} are observable as competing signals over a finite range of filling factor ν\nu and are indistinguishable unless polarized light is used.

IV renormalization

The self-energies Δϵn\Delta\epsilon_{n} are afflicted with ultraviolet divergences. In this section we consider how to extract physically observable information out of them. To this end, one first defines renormalized parameters by setting v=Zvvren=vren+δvv=Z_{v}v^{\rm ren}=v^{\rm ren}+\delta v, γ1=γ1ren+δγ1\gamma_{1}=\gamma_{1}^{\rm ren}+\delta\gamma_{1}, Δ=Δren+δΔ\Delta=\Delta^{\rm ren}+\delta\Delta, etc., with counterterms δv=(Zv1)vrenO(αe)\delta v=(Z_{v}-1)\,v^{\rm ren}\sim O(\alpha_{e}), δγ1\delta\gamma_{1}, etc. The one-body Hamiltonian H=Hren+δctHH=H^{\rm ren}+\delta_{\rm ct}H is then divided into the portion HrenH^{\rm ren} involving only the renormalized quantities and the counterterms δctH\delta_{\rm ct}H. One starts with HrenH^{\rm ren}, calculates the Coulombic corrections and encounters divergences. If one could remove them by adjusting δctH\delta_{\rm ct}H, renormalization is done properly.

Fortunately, the renormalization procedure to O(V)O(αe)O(V)\sim O(\alpha_{e}) for bilayer graphene in a magnetic field BB has been formulated earlier KS_CR_eh . The key point is that, since BB simply acts as a long-wavelength cutoff \sim\ell, the short-distance structure of the present theory is known from its free-space (B=0B=0) version: (i) The divergence associated with velocity vv is the same as that for monolayer graphene,

δv=(αe/8ϵb)log(Λ)2,\delta v=-(\alpha_{e}/8\epsilon_{b})\log(\ell\Lambda)^{2}, (44)

with a momentum cutoff Λ\Lambda. (ii) v4v_{4} and uu remain finite,

δv4=δu=0.\delta v_{4}=\delta u=0. (45)

We thus take r=v4/vrenr=v_{4}/v^{\rm ren} and μ=(u/2)/ωcren\mu=(u/2)/\omega_{c}^{\rm ren} finite; ωcren2vren/\omega_{c}^{\rm ren}\equiv\sqrt{2}v^{\rm ren}/\ell. (iii) γ1\gamma_{1} and Δ\Delta, though mixed under renormalization, are also governed by δv\delta v,

δγ1\displaystyle\delta\gamma_{1} =\displaystyle= (γ1ren+rΔren)h(r)δv/vren,\displaystyle(\gamma_{1}^{\rm ren}+r\,\Delta^{\rm ren})\,h(r)\,\delta v/v^{\rm ren},
δΔ\displaystyle\delta\Delta =\displaystyle= 2(Δren+rγ1ren)h(r)δv/vren,\displaystyle 2\,(\Delta^{\rm ren}+r\,\gamma_{1}^{\rm ren})\,h(r)\,\delta v/v^{\rm ren}, (46)

where h(r)=1/(1r2)h(r)=1/(1-r^{2}).

Let us now rewrite the bare level spectra ϵn=ωcen(g;μ,d,r)\epsilon_{n}=\omega_{c}\,e_{n}(g;\mu,d,r) as ϵn=ϵnren+δϵn\epsilon_{n}=\epsilon_{n}^{\rm ren}+\delta\epsilon_{n}. The renormalized spectra ϵnren\epsilon_{n}^{\rm ren} are the portion consisting of ωcren\omega_{c}^{\rm ren}, grenγ1ren/ωcreng^{\rm ren}\equiv\gamma_{1}^{\rm ren}/\omega_{c}^{\rm ren}, drenΔren/ωcrend^{\rm ren}\equiv\Delta^{\rm ren}/\omega_{c}^{\rm ren}, etc. The associated counterterms δϵn\delta\epsilon_{n}, to be written as δϵnλnδv/vren\delta\epsilon_{n}\equiv\lambda_{n}\,\delta v/v^{\rm ren} with finite coefficients λn\lambda_{n}, are expressed in terms of (δv,δγ1,δΔ)(\delta v,\delta\gamma_{1},\delta\Delta) and are uniquely fixed once one specifies vrenv^{\rm ren} (or, the divergence δv\delta v) by referring to a specific observable quantity; see Appendix A for details. The dressed level spectra, rewritten as

ϵ^n=ϵn+Δϵn=ϵnren+(Δϵn)ren,\hat{\epsilon}_{n}=\epsilon_{n}+\Delta\epsilon_{n}=\epsilon_{n}^{\rm ren}+(\Delta\epsilon_{n})^{\rm ren}, (47)

thereby reveal observable self-energies (Δϵn)renΔϵn+δϵn(\Delta\epsilon_{n})^{\rm ren}\equiv\Delta\epsilon_{n}+\delta\epsilon_{n}, with the divergences in Δϵn\Delta\epsilon_{n} removed by δϵn\delta\epsilon_{n}. Also the full CR energies ϵexcnj\epsilon_{\rm exc}^{n\leftarrow j} in Eq. (42) take a renormalized form,

ϵexcnj=ϵnrenϵjren+(Δϵn,j)ren,\epsilon_{\rm exc}^{n\leftarrow j}=\epsilon_{n}^{\rm ren}-\epsilon_{j}^{\rm ren}+(\Delta\epsilon^{n,j})^{\rm ren}, (48)

with corrections (Δϵn,j)ren(\Delta\epsilon^{n,j})^{\rm ren} free from divergences.

Let us here define vrenv^{\rm ren} by referring to CR in the 23-2\leftarrow-3 channel, as chosen experimentally HJTS . We fix vrenv^{\rm ren} so that ϵexc23|(μ,d,r)0nf=3\epsilon^{-2\leftarrow-3}_{\rm exc}|^{n_{\rm f}=-3}_{(\mu,d,r)\rightarrow 0} takes a naive form (ϵ2renϵ3ren)|X=0(\epsilon_{-2}^{\rm ren}-\epsilon_{-3}^{\rm ren})|_{X=0}; i.e., (Δϵ2,3)ren|X=0=0(\Delta\epsilon^{-2,-3})^{\rm ren}|_{X=0}=0. A neat way to handle this prescription is to replace n(z){\cal F}_{n}(z) in Δϵn\Delta\epsilon_{n} [of Eq. (36)] by

nren(z)=n(z)2λn[Γ23(z)]X=0,{\cal F}^{\rm ren}_{n}(z)={\cal F}_{n}(z)-2\lambda_{n}\,[\Gamma^{-2\leftarrow-3}(z)]_{X=0}, (49)

and cast (Δϵn)ren(\Delta\epsilon_{n})^{\rm ren} in the form

(Δϵn)ren\displaystyle(\Delta\epsilon_{n})^{\rm ren} =\displaystyle= Ωnkν[k]𝐩v𝐩γ𝐩2|g𝐩nk|2,\displaystyle\Omega_{n}-\sum_{k}\nu[k]\,\sum_{\bf p}v_{\bf p}\gamma_{\bf p}^{2}|g^{nk}_{\bf p}|^{2}, (50)
Ωn\displaystyle\Omega_{n} =\displaystyle= 12𝐩v𝐩nren(z).\displaystyle{\textstyle{1\over{2}}}\sum_{\bf p}v_{\bf p}\,{\cal F}_{n}^{\rm ren}(z). (51)

Here Γ23(z)=γ𝐩2Iz/(λ2λ3)\Gamma^{-2\leftarrow-3}(z)=\gamma^{2}_{\bf p}\,I_{z}/(\lambda_{-2}-\lambda_{-3}), with

Iz=k3(|g𝐩2,k|2|g𝐩3,k|2)g𝐩2,2g𝐩3,3,I_{z}=-\sum_{k\leq-3}(|g^{-2,k}_{\bf-p}|^{2}-|g^{-3,k}_{\bf p}|^{2})-g^{-2,-2}_{\bf p}\,g^{-3,-3}_{\bf-p}, (52)

is a kernel associated with Δϵ2,3\Delta\epsilon^{-2,-3}. In this renormalized form, nren(z){\cal F}_{n}^{\rm ren}(z) are sizable only for |𝐩|O(1)\ell|{\bf p}|\sim O(1) and vanish rapidly as z=122𝐩2z={1\over{2}}\ell^{2}{\bf p}^{2}\rightarrow\infty; one can thus calculate many-body corrections Ωn\Omega_{n} numerically without handling divergences. Note that we have carried out renormalization in such a way that nren(z){\cal F}^{\rm ren}_{n}(z) and hence Ωn\Omega_{n} also share the same conjugation properties as Fn(z)F_{n}(z) in Eq. (35), i.e.,

Ωn=Ωn|X,ΩnK=ΩnK|μ=ΩnK|d,r.\Omega_{-n}=-\Omega_{n}|_{-X},\ \ \Omega_{n}^{K^{\prime}}=\Omega_{n}^{K}|_{-\mu}=-\Omega_{-n}^{K}|_{-d,-r}. (53)

Once the eigenmodes 𝐛n{\bf b}_{n} are fixed numerically for a given set of parameters (v,γ1,u,Δ,r)(v,\gamma_{1},u,\Delta,r), one can now construct g𝐩mng^{mn}_{\bf p}, (ϵn,λn)(\epsilon_{n},\lambda_{n}), and nren(z){\cal F}^{\rm ren}_{n}(z) and calculate the spectra ϵ^n\hat{\epsilon}_{n} and ϵexcnj\epsilon_{\rm exc}^{n\leftarrow j}. Table I shows a list of Ωn\Omega_{n} of our interest, in units of

V~c=𝐩v𝐩γ𝐩2=αϵbπ270.3ϵbB[T]meV.\tilde{V}_{c}=\sum_{\bf p}v_{\bf p}\gamma_{\bf p}^{2}={\alpha\over{\epsilon_{b}\ell}}\sqrt{{\pi\over{2}}}\approx{70.3\over{\epsilon_{b}}}\,\sqrt{B[{\rm T}]}\,{\rm meV}. (54)

For the PZM levels, Ω0\Omega_{0} and Ω1\Omega_{1} are practically linear in μ\mu and are barely modified by ee-hh breaking (d,r)(d,r); indeed, (Ω0,Ω1)|d=r=0(0.6457,0.6120)μV~c(\Omega_{0},\Omega_{1})|_{d=r=0}\approx(-0.6457,-0.6120)\,\mu\,\tilde{V}_{c}. For other levels ee-hh breaking is evident, Ωn=Ωn|XΩn\Omega_{n}=-\Omega_{-n}|_{-X}\not=-\Omega_{-n}.

Table 1: Many-body corrections Ωn\Omega_{n} in valley KK at B=20B=20T. Setting μμ\mu\rightarrow-\mu yields Ωn\Omega_{n} in valley KK^{\prime}.
Ω4V~c(0.4758+0.0635μ0.0299μ2)\Omega_{4}\approx\tilde{V}_{c}\,(0.4758+0.0635\mu-0.0299\mu^{2})
Ω3V~c(0.4363+0.0884μ+0.157μ2)\Omega_{3}\approx\tilde{V}_{c}\,(0.4363+0.0884\mu+0.157\mu^{2})
Ω2V~c(0.3539+0.146μ+0.663μ2)\Omega_{2}\approx\tilde{V}_{c}\,(0.3539+0.146\mu+0.663\mu^{2})
Ω0V~c( 00.6446μ0.0170μ2)\Omega_{0}\approx\tilde{V}_{c}\,(\ \ 0\ \ \ \ \ -0.6446\mu-0.0170\mu^{2})
Ω1V~c(0.003390.6109μ0.0317μ2)\Omega_{1}\approx\tilde{V}_{c}\,(-0.00339-0.6109\mu-0.0317\mu^{2})
Ω2V~c(0.4136+0.222μ0.641μ2)\Omega_{-2}\!\approx\tilde{V}_{c}\,(-0.4136+0.222\mu-0.641\mu^{2})
Ω3V~c(0.4995+0.114μ0.166μ2)\Omega_{-3}\!\approx\tilde{V}_{c}\,(-0.4995+0.114\mu-0.166\mu^{2})
Ω4V~c(0.5441+0.0719μ+0.00489μ2)\Omega_{-4}\!\approx\tilde{V}_{c}\,(-0.5441+0.0719\mu+0.00489\mu^{2})

V Orbital mixing and Cyclotron resonance within the PZM sector

In bilayer graphene the LLL consists of 2spin×2valley=42_{\rm spin}\times 2_{\rm valley}=4 sets of nearly degenerate n={0,1}n=\{0,1\} levels. For the empty PZM sector [of a given (spin, valley)] at Nf=0N_{\rm f}=0 one can explicitly write down the renormalized spectra by substituting ν[k]=12(δk,0+δk,1)\nu[k]=-{1\over{2}}(\delta_{k,0}+\delta_{k,1}) into Eq. (50),

ϵ^0\displaystyle\hat{\epsilon}_{0}\!\!\! =Nf=0\displaystyle\stackrel{{\scriptstyle N_{\rm f}=0}}{{=}} ωcμ+Ω0+12(1+12c12)V~c,\displaystyle\!\!\!-\omega_{c}\mu+\Omega_{0}+{\textstyle{1\over{2}}}(1+{\textstyle{1\over{2}}}\,c_{1}^{2})\,\tilde{V}_{c},
ϵ^1\displaystyle\hat{\epsilon}_{1}\!\!\! =Nf=0\displaystyle\stackrel{{\scriptstyle N_{\rm f}=0}}{{=}} ωc(μκ)+Ω1+12(1+12c1214C)V~c,\displaystyle\!\!\!-\omega_{c}(\mu-\kappa)+\Omega_{1}+{\textstyle{1\over{2}}}(1+{\textstyle{1\over{2}}}\,c_{1}^{2}-{\textstyle{1\over{4}}}\,C)\tilde{V}_{c}, (55)

with C=(43c12)c12C=(4-3c_{1}^{2})\,c_{1}^{2}. Here and from now on all quantities refer to renormalized ones. In this section we use ϵ^n\hat{\epsilon}_{n} to denote the Nf=0N_{\rm f}=0 spectra, ϵ^n=ϵ^n|Nf=0\hat{\epsilon}_{n}=\hat{\epsilon}_{n}|^{N_{\rm f}=0}.

In Eq. (55), the first terms (μ,κ,Ω0,Ω1)\ni(\mu,\kappa,\Omega_{0},\Omega_{1}) are odd in breaking X=(μ,d,r)X=(\mu,d,r) while the last terms V~c\propto\tilde{V}_{c} are even in XX. For X=0X=0, only the latter remain. The orbital degeneracy is therefore lifted by quantum corrections even for zero breaking X=0X=0, giving rise to the orbital Lamb shift KS_Ls , with ϵ^1\hat{\epsilon}_{1} lower than ϵ^0\hat{\epsilon}_{0} by

ϵLs18CV~c.\epsilon_{\rm Ls}\equiv{\textstyle{1\over{8}}}\,C\,\tilde{V}_{c}. (56)

Numerically, at B=20B=20 T, C1.21+0.016μ+0.04μ2C\approx 1.21+0.016\mu+0.04\mu^{2}, (ϵ^0,ϵ^1)|X=0(0.72,0.57)V~c(\hat{\epsilon}_{0},\hat{\epsilon}_{1})|_{X=0}\approx(0.72,0.57)\tilde{V}_{c} and ϵLs0.15V~c\epsilon_{\rm Ls}\approx 0.15\tilde{V}_{c}, or ϵLs8.3\epsilon_{\rm Ls}\approx 8.3 meV for a choice of V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 or ϵb5.7\epsilon_{b}\approx 5.7. The dressed spectra ϵ^n\hat{\epsilon}_{n} considerably deviate from the one-body spectra ϵn\epsilon_{n}, as shown in Fig. 1 for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 and B=20B=20T at bias u=10u=10 meV.

Refer to caption
Figure 1: Landau-level spectra in valleys (K,K)(K,K^{\prime}) at bias u=u= 10 meV and B=20B=20 T; spin splitting is suppressed. (left) One-body spectra ϵn|d=r=0\epsilon_{n}|_{d=r=0} and ϵn\epsilon_{n}. ee-hh breaking (d,r)(d,r) shifts the ee-hh symmetric spectra ϵn|d=r=0\epsilon_{n}|_{d=r=0} upwards appreciably, except for ϵ0\epsilon_{0}. (right) Coulomb-corrected spectra ϵ^n\hat{\epsilon}_{n} for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4. Level spectra evolve with filling (NfK,NfK)(N_{\rm f}^{K},N_{\rm f}^{K^{\prime}}) of each valley and spin. When bias uu dominates over spin splitting μZ\mu_{\rm Z}, filling of the LLL will start with valley KK with spin \downarrow at ν=4\nu=-4, as illustrated in the figure. Green arrows denote possible ν=0\nu=0 gaps, purple arrows ν=2\nu=\mp 2 gaps and red arrows orbital gaps at odd-integer filling.

Let us write, for X0X\not=0, the full (0,1)(0,1) shift as

ϵ^0ϵ^1=ϵLs+Ω0Ω1ωcκ(1ξ)ϵLs.\hat{\epsilon}_{0}-\hat{\epsilon}_{1}=\epsilon_{\rm Ls}+\Omega_{0}-\Omega_{1}-\omega_{c}\kappa\equiv(1-\xi)\,\epsilon_{\rm Ls}. (57)

Numerically, at B=20B=20T,

ξ0.023+0.22μ+(0.325+1.7μ)/(V~c/ωc).\xi\approx-0.023+0.22\mu+(0.325+1.7\mu)/(\tilde{V}_{c}/\omega_{c}). (58)

Obviously, ξ>0\xi>0 for μ0\mu\geq 0. The O(X)O(X) term in ϵ^1\hat{\epsilon}_{1}, κ2μ/(g2+1)+O(d,r)\kappa\approx 2\mu/(g^{2}+1)+O(d,r), thus tends to reduce the orbital Lamb shift, e.g., (1ξ)ϵLs(0.53,0.21,0.11)ϵLs(1-\xi)\,\epsilon_{\rm Ls}\approx(0.53,0.21,-0.11)\epsilon_{\rm Ls} for u=(20,0,20)u=(-20,0,20) meV at B=20B=20T. At the same time, κ\kappa makes ϵ^1\hat{\epsilon}_{1} slightly less dependent on μ\mu than ϵ^0\hat{\epsilon}_{0}. As a result, the nearly degenerate PZM levels (ϵ^0,ϵ^1)(\hat{\epsilon}_{0},\hat{\epsilon}_{1}), being quasi linear in μ\mu, necessarily have a crossing or level inversion in either valley (per spin) as bias uu is varied.

Refer to caption
Figure 2: Level spectra at B=B=20 T for (a) V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 and (b) V~c/ωc=0.2\tilde{V}_{c}/\omega_{c}=0.2. Bold (solid/long-dashed) curves refer to the spectra (in valley K/KK/K^{\prime}) at Nf=0N_{\rm f}=0 [with empty n=(0,1)n=(0,1) levels], and thin (dashed/dotted) curves to those at Nf=2N_{\rm f}=2 [with filled (0,1)(0,1) levels]. Level inversion takes place at Nf=0N_{\rm f}=0 across u=ucru=u^{\rm cr}; ucru^{\rm cr}\approx13 meV in (a) and ucru^{\rm cr}\approx -19 meV in (b). The ν=0\nu=0 band gaps (green arrows) increase with bias uu while possible ν=2\nu=\mp 2 gaps (purple arrows) barely change. (c) ucru^{\rm cr} as a function of V~c/ωc\tilde{V}_{c}/\omega_{c} at B=B=(10,20,30) T.

These features are indeed seen from Fig. 2(a), which shows how such Nf=0N_{\rm f}=0 level spectra (per spin) change with bias uu for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4; solid curves refer to valley KK and long-dashed curves to KK^{\prime}; ϵ^n|K=ϵ^n|μK\hat{\epsilon}_{n}|^{K^{\prime}}=\hat{\epsilon}_{n}|^{K}_{-\mu}. Level inversion takes place across ξ=1\xi=1, or numerically, across μcr0.0475\mu^{\rm cr}\approx 0.0475 or ucr13u^{\rm cr}\approx 13 meV. Such a critical bias ucru^{\rm cr} gets smaller as V~c\tilde{V}_{c} is made weaker and for higher BB, as seen from Figs. 2(b) and 2(c); ucr>0u^{\rm cr}>0 lies in valley KK while ucr<0u^{\rm cr}<0 lies in valley KK^{\prime}. Level inversion is also present for d=r=0d=r=0, with ucr62u^{\rm cr}\approx 62meV at B=20B=20T.

When the PZM sector is gradually filled with electrons, those level spectra come down (via exchange interaction) and, when the sector is filled up, they turn into the ones, depicted with thin dashed and dotted curves in Fig. 2. These Nf=2N_{\rm f}=2 spectra are ee-hh conjugate to the Nf=0N_{\rm f}=0 spectra,

(ϵ^0,ϵ^1)|Nf=2=(ϵ^0,ϵ^1)|XNf=0,(\hat{\epsilon}_{0},\hat{\epsilon}_{1})|^{N_{\rm f}=2}=(-\hat{\epsilon}_{0},-\hat{\epsilon}_{1})|^{N_{\rm f}=0}_{-X}, (59)

and are given by

ϵ^0|Nf=2\displaystyle\hat{\epsilon}_{0}|^{N_{\rm f}=2}\! =\displaystyle= ωcμ+Ω012(1+12c12)V~c,\displaystyle\!-\omega_{c}\mu+\Omega_{0}-{\textstyle{1\over{2}}}(1+{\textstyle{1\over{2}}}\,c_{1}^{2})\,\tilde{V}_{c},
ϵ^1|Nf=2\displaystyle\hat{\epsilon}_{1}|^{N_{\rm f}=2}\! =\displaystyle= ωc(μκ)+Ω112(1+12c1214C)V~c.\displaystyle\!-\omega_{c}(\mu-\kappa)+\Omega_{1}-{\textstyle{1\over{2}}}(1+{\textstyle{1\over{2}}}c_{1}^{2}-{\textstyle{1\over{4}}}C)\tilde{V}_{c}.\ \ (60)

At Nf=2N_{\rm f}=2, orbital shift (ϵ^0ϵ^1)|Nf=2=(1+ξ)ϵLs<0(\hat{\epsilon}_{0}-\hat{\epsilon}_{1})|^{N_{\rm f}=2}=-(1+\xi)\,\epsilon_{\rm Ls}<0 is reversed in the (0,1)(0,1) layout and is considerably enhanced by ξ\xi, as is clear from Fig. 2. Incidentally, for (d,r)=0(d,r)=0, these conjugate Nf=(0,2)N_{\rm f}=(0,2) spectra look symmetric about the uu axis. In this sense, ee-hh breaking (d,r)(d,r) is seen as an apparent asymmetry in Fig. 2. [Actually, asymmetry is particularly sizable for the n=1n=1 level; with Rnϵ^n/(ϵ^n|d=r=0)R_{n}\equiv\hat{\epsilon}_{n}/(\hat{\epsilon}_{n}|_{d=r=0}), one finds R11.2R_{1}\approx 1.2, R01R_{0}\approx 1 and Rn1.07R_{n}\approx 1.07 (n=±2±6n=\pm 2\sim\pm 6) for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 and u0u\sim 0.]

Let us now consider what will happen when we pass from Nf=0N_{\rm f}=0 to Nf=2N_{\rm f}=2 in valley KK with bias uu kept fixed in the range 0<u<ucr0<u<u^{\rm cr}. It is the n=1n=1 level that starts to be filled, getting lower in energy. The n=0n=0 level also follows but, when Nf=2N_{\rm f}=2 is reached, it gets even lower than the n=1n=1 level. This signals a level crossing or instability with filling, which actually is avoided via mixing of n={0,1}n=\{0,1\} levels KS_Ls . For u>ucru>u^{\rm cr} no such mixing takes place. Orbital mixing is therefore triggered by level inversion in the PZM spectra. (i) When ucr0u^{\rm cr}\geq 0, orbital mixing arises in both valleys for 0u<ucr0\leq u<u^{\rm cr} and in one valley (KK^{\prime}) for u>ucru>u^{\rm cr}. (ii) When ucr<0u^{\rm cr}<0, mixing takes place only in one valley (KK^{\prime}) for u>|ucr|u>|u^{\rm cr}|, as seen from Fig. 2(b). (Here we suppose no level inversion in the Nf=2N_{\rm f}=2 spectra and take, e.g., u50u\lesssim 50 meV.)

To see how level mixing proceeds let us rotate ψ^=(ψ0,ψ1)t\hat{\psi}=(\psi^{0},\psi^{1})^{\rm t} to Φ^=(Φ0,Φ1)t=Uψ^\hat{\Phi}=(\Phi^{0},\Phi^{1})^{\rm t}=U\hat{\psi} in the orbital space,

ψ^=UΦ^=(cθsθsθcθ)(Φ0Φ1),\hat{\psi}=U\hat{\Phi}=\left(\begin{array}[]{cc}c_{\theta}&-s_{\theta}\\ s_{\theta}&c_{\theta}\\ \end{array}\right)\left(\begin{array}[]{c}\Phi^{0}\\ \Phi^{1}\\ \end{array}\right), (61)

where cθcos(θ/2)c_{\theta}\equiv\cos(\theta/2) and sθsin(θ/2)s_{\theta}\equiv\sin(\theta/2); we refer to the levels associated with Φn\Phi^{n} as n=(0θ,1θ)n=(0_{\theta},1_{\theta}). We also define the rotated charges S𝐩mnS_{\bf p}^{mn} by g𝐩mnR𝐩mn=G𝐩mnS𝐩mng^{mn}_{\bf-p}\,R^{mn}_{\bf p}=G^{mn}_{\bf-p}\,S_{\bf p}^{mn}, where G𝐩mn=(Ug𝐩U)mnG^{mn}_{\bf-p}=(U^{{\dagger}}g_{\bf-p}U)^{mn} for n,m(0,1)n,m\in(0,1), G𝐩mn=(Ug𝐩)mnG^{mn}_{\bf-p}=(U^{{\dagger}}g_{\bf-p})^{mn} for m(0,1)m\in(0,1) and n(0,1)n\notin(0,1), etc.; S𝐩mnS_{\bf-p}^{mn} are written with (Φ0,Φ1)(\Phi^{0},\Phi^{1}) and ψj\psi^{j} with j(0,1)j\notin(0,1).

The PZM sector for 0Nf20\leq N_{\rm f}\leq 2 is now governed by the one-body Hamiltonian [with the Nf=0N_{\rm f}=0 spectra (ϵ^0,ϵ^1)(\hat{\epsilon}_{0},\hat{\epsilon}_{1}) for (ψ0,ψ1)(\psi^{0},\psi^{1})] plus Coulomb interaction VpzmV^{\rm pzm} [with g𝐩R𝐩G𝐩S𝐩g_{\bf p}R_{\bf-p}\rightarrow G_{\bf p}S_{\bf-p} in Eq. (28)] acting within this sector. One can readily diagonalize it using the Hartree-Fock approximation: The rotated levels (0θ,1θ)(0_{\theta},1_{\theta}) have the spectra KS_Ls ,

ϵ^0θ\displaystyle\hat{\epsilon}_{0_{\theta}} =\displaystyle= ϵ^++ϵ^cosθ+N0E00(θ)+N1E01(θ),\displaystyle\hat{\epsilon}_{+}+\hat{\epsilon}_{-}\cos\theta+N_{0}\,E^{00}(\theta)+N_{1}\,E^{01}(\theta),
ϵ^1θ\displaystyle\hat{\epsilon}_{1_{\theta}} =\displaystyle= ϵ^+ϵ^cosθ+N1E11(θ)+N0E01(θ),\displaystyle\hat{\epsilon}_{+}-\hat{\epsilon}_{-}\cos\theta+N_{1}\,E^{11}(\theta)+N_{0}\,E^{01}(\theta), (62)

where ϵ^+=12(ϵ^0+ϵ^1)\hat{\epsilon}_{+}={1\over{2}}\,(\hat{\epsilon}_{0}+\hat{\epsilon}_{1}), ϵ^=12(1ξ)ϵLs\hat{\epsilon}_{-}={1\over{2}}\,(1-\xi)\,\epsilon_{\rm Ls} and

E00(θ)\displaystyle E^{00}(\theta) =\displaystyle= V~c[1116C(1cosθ)2],\displaystyle-\tilde{V}_{c}\,\left[1-\textstyle{1\over{16}}C\,(1-\cos\theta)^{2}\right],
E11(θ)\displaystyle E^{11}(\theta) =\displaystyle= V~c[1116C(1+cosθ)2],\displaystyle-\tilde{V}_{c}\,\left[1-\textstyle{1\over{16}}C\,(1+\cos\theta)^{2}\right],
E01(θ)\displaystyle E^{01}(\theta) =\displaystyle= V~c[12c12116C{1(cosθ)2}];\displaystyle-\tilde{V}_{c}\,\left[\textstyle{1\over{2}}c_{1}^{2}-{1\over{16}}\,C\{1-(\cos\theta)^{2}\}\right]; (63)

(N0,N1)(N_{0},N_{1}) denote the filling fractions of the (0θ,1θ)(0_{\theta},1_{\theta}) levels, with 0N010\leq N_{0}\leq 1, etc. Note that (ϵ^0θ,ϵ^1θ)(\hat{\epsilon}_{0_{\theta}},\hat{\epsilon}_{1_{\theta}}) enjoy the reciprocal relation

ϵ^1θ=ϵ^0θ|N0N1,θπθ.\hat{\epsilon}_{1_{\theta}}=\hat{\epsilon}_{0_{\theta}}|_{N_{0}\leftrightarrow N_{1},\theta\rightarrow\pi-\theta}. (64)

Diagonalization of the PZM spectra is achieved for θ\theta that obeys the relation

θ[ϵ0(θ)+12{N0E00(θ)N1E11(θ)}]\displaystyle\partial_{\theta}\left[\epsilon_{0}(\theta)+\textstyle{1\over{2}}\,\{N_{0}\,E^{00}(\theta)-N_{1}\,{E}^{11}(\theta)\}\right] (65)
\displaystyle\propto sinθ[N0+N1(1ξ)+(N1N0)cosθ]=0,\displaystyle\sin\theta\left[N_{0}+N_{1}-(1-\xi)+(N_{1}-N_{0})\cos\theta\right]=0,\ \ \ \ \

i.e., sinθ=0\sin\theta=0 or cosθ=(1ξ+N0N1)/(N1N0)\cos\theta=(1-\xi+N_{0}-N_{1})/(N_{1}-N_{0}), where ϵ0(θ)=cθ2ϵ^0+sθ2ϵ^1=ϵ^++ϵ^cosθ\epsilon_{0}(\theta)=c_{\theta}^{2}\,\hat{\epsilon}_{0}+s_{\theta}^{2}\,\hat{\epsilon}_{1}=\hat{\epsilon}_{+}+\hat{\epsilon}_{-}\cos\theta.

Refer to caption
Figure 3: Orbital mixing. (a) PZM spectra (ϵ^0θ,ϵ^1θ)(\hat{\epsilon}_{0_{\theta}},\hat{\epsilon}_{1_{\theta}}) over the range 0Nf20\leq N_{\rm f}\leq 2 for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 at u=0u=0 and B=20B=20 T. Dashed lines refer to the level spectra with no rotation θ=0\theta=0. (b) Cyclotron resonance within the PZM sector at bias u=(0,±20)u=(0,\pm 20) meV.

Figure 3(a) depicts how the (ϵ^0θ,ϵ^1θ)(\hat{\epsilon}_{0_{\theta}},\hat{\epsilon}_{1_{\theta}}) spectra change with filling factor Nf=N1+N0N_{\rm f}=N_{1}+N_{0} for u=0<ucru=0<u^{\rm cr} and B=20B=20 T. For u<ucru<u^{\rm cr} critical changes arise at three points (Nf,1,Nf+=1+Nf)(N_{\rm f}^{-},1,N_{\rm f}^{+}=1+N_{\rm f}^{-}) with Nf=12(1ξ)<0.5N_{\rm f}^{-}={1\over{2}}(1-\xi)<0.5, and orbital mixing takes place in the range (Nf,Nf+)(N_{\rm f}^{-},N_{\rm f}^{+}): (i) The 1θ1_{\theta} level first gets filled. θ\theta deviates from zero at Nf(=N1)=NfN_{\rm f}(=N_{1})=N_{\rm f}^{-}, and changes as

cosθ=(1ξ)/Nf1,\cos\theta=(1-\xi)/N_{\rm f}-1, (66)

until cosθ=ξ<0\cos\theta=-\xi<0 is reached at Nf=1N_{\rm f}=1. (ii) Filling of the 0θ0_{\theta} level then starts, and θ\theta changes according to

cosθ=(1ξNf)/(2Nf).\cos\theta=(1-\xi-N_{\rm f})/(2-N_{\rm f}). (67)

cosθ\cos\theta attains 1-1 or θ=π\theta=\pi at Nf=Nf+N_{\rm f}=N_{\rm f}^{+} and ceases to change. Via mixing the {0,1}\{0,1\} levels are interchanged,

(ψ0,ψ1))|θ=0(Φ0,Φ1)|θ=π=(ψ1,ψ0).(\psi^{0},\psi^{1}))|_{\theta=0}\rightarrow(\Phi^{0},\Phi^{1})|_{\theta=\pi}=(\psi^{1},-\psi^{0}). (68)

Actually, noting Eq. (65), one can simplify the level spectra when θ\theta is moving, i.e., for NfNfNf+N_{\rm f}^{-}\leq N_{\rm f}\leq N_{\rm f}^{+},

ϵ^0θ\displaystyle\hat{\epsilon}_{0_{\theta}} =\displaystyle= ϵ^0V~c(N0+12c12N1),\displaystyle\hat{\epsilon}_{0}-\tilde{V}_{c}\,(N_{0}+{\textstyle{1\over{2}}}c_{1}^{2}N_{1}),
ϵ^1θ\displaystyle\hat{\epsilon}_{1_{\theta}} =\displaystyle= ϵ^0V~c(N1+12c12N0).\displaystyle\textstyle\hat{\epsilon}_{0}-\tilde{V}_{c}\,(N_{1}+{\textstyle{1\over{2}}}c_{1}^{2}N_{0}). (69)

The Nf=1N_{\rm f}=1 spectra (ϵ^0θ,ϵ^1θ)|Nf=1=(ϵ^012c12V~c,ϵ^0V~c)(\hat{\epsilon}_{0_{\theta}},\hat{\epsilon}_{1_{\theta}})|^{N_{\rm f}=1}=(\hat{\epsilon}_{0}-{\textstyle{1\over{2}}}c_{1}^{2}\tilde{V}_{c},\hat{\epsilon}_{0}-\tilde{V}_{c}) then obey the relation

(ϵ^0θ,ϵ^1θ)|Nf=1=(ϵ^1θ,ϵ^0θ)|XNf=1,(\hat{\epsilon}_{0_{\theta}},\hat{\epsilon}_{1_{\theta}})|^{N_{\rm f}=1}=(-\hat{\epsilon}_{1_{\theta}},-\hat{\epsilon}_{0_{\theta}})|^{N_{\rm f}=1}_{-X}, (70)

which implies that, via ee-hh conjugation, the filled 1θ1_{\theta} level turns into the filled 0θ0_{\theta} within a valley. The small gap ϵ^0ϵ^1=(1ξ)ϵLs\hat{\epsilon}_{0}-\hat{\epsilon}_{1}=(1-\xi)\,\epsilon_{\rm Ls} at Nf=0N_{\rm f}=0 develops into a sizable orbital gap (112c12)V~c(1-{\textstyle{1\over{2}}}c_{1}^{2})\,\tilde{V}_{c} at Nf=1N_{\rm f}=1, as seen from Fig. 3(a).

On the other hand, for u>ucru>u^{\rm cr}, no such mixing arises, but (ϵ^0,ϵ^1)(\hat{\epsilon}_{0},\hat{\epsilon}_{1}) show similar behavior with filling. At Nf=1N_{\rm f}=1, (ϵ^0,ϵ^1)|Nf=1=(ϵ^0V~c,ϵ^112c12V~c)(\hat{\epsilon}_{0},\hat{\epsilon}_{1})|^{N_{\rm f}=1}=(\hat{\epsilon}_{0}-\tilde{V}_{c},\hat{\epsilon}_{1}-{\textstyle{1\over{2}}}c_{1}^{2}\,\tilde{V}_{c}) correctly obey the relation in Eq. (40) and have again a sizable gap.

In bilayer graphene the empty LLL (at ν=4\nu=-4) consists of four sets of empty PZM sectors. Let us now consider how the LLL is filled over the range 4ν4-4\leq\nu\leq 4. As an illustration we focus on some typical cases in which the eightfold degeneracy of the LLL is fully lifted. Suppose first that bias uu is chosen so that the valley gap O(u)\sim O(u) dominates over the spin gap μZ\mu_{\rm Z}. Filling of the LLL will then starts in valley KK with a spin component of the lowest energy, and each {valley, spin} set of PZM sectors will be filled in the following order

ν=4{K}ν=2{K}ν=0{K}ν=2{K}.\stackrel{{\scriptstyle\nu=-4}}{{\rightarrow}}\{K\downarrow\}\stackrel{{\scriptstyle\nu=-2}}{{\rightarrow}}\{K\uparrow\}\stackrel{{\scriptstyle\nu=0}}{{\rightarrow}}\{K^{\prime}\downarrow\}\stackrel{{\scriptstyle\nu=2}}{{\rightarrow}}\{K^{\prime}\uparrow\}. (71)

Figure 1 illustrates such a sequence of level spectra {ϵ^n}\{\hat{\epsilon}_{n}\} for u=10meV<ucru=10{\rm meV}<u^{\rm cr}. There the total filling factor ν\nu increases with each valley-filling factor {NfK,NfK,}\{N_{\rm f}^{K\downarrow},N_{\rm f}^{K\uparrow},\cdots\}, and a sizable band gap emerges at each integer filling: (i) The ν=±3\nu=\pm 3 and ν=±1\nu=\pm 1 gaps are orbital gaps (in red) of width (112c12)V~c(0.56V~c)(1-{1\over{2}}c_{1}^{2})\,\tilde{V}_{c}\ (\sim 0.56\tilde{V}_{c}). (ii) The ν=2\nu=\mp 2 gaps are enhanced spin gaps (in purple), ϵ^1|μNf=0ϵ^0|Nf=2(1+12c1214C)V~c(1.13V~c)\hat{\epsilon}_{1}|^{N_{\rm f}=0}_{-\mu}-\hat{\epsilon}_{0}|^{N_{\rm f}=2}\approx(1+{1\over{2}}c_{1}^{2}-{1\over{4}}C)\,\tilde{V}_{c}\ (\sim 1.13\tilde{V}_{c}), associated with full filling of one valley. (iii) Most prominent is the ν=0\nu=0 gap, which is a valley (+ orbital) gap (in green), ϵ^1|μK;Nf=0ϵ^1|K;Nf=2\hat{\epsilon}_{1}|^{K^{\prime};N_{\rm f}=0}_{-\mu}-\hat{\epsilon}_{1}|^{K;N_{\rm f}=2} u+(1+12c1214C)V~c\sim u+(1+{1\over{2}}c_{1}^{2}-{1\over{4}}C)\,\tilde{V}_{c}. As seen from Figs. 2(a) and 2(b), the ν=0\nu=0 gap and ν=±2\nu=\pm 2 gaps are essentially the same at zero bias u=0u=0, but the former increases rapidly with uu while the latter barely change.

On the other hand, when the spin gap dominates over the valley gap in high field BB, i.e., μZu+0\mu_{Z}\gg u\sim+0, filling of the LLL will proceed in the following order

ν=4{K}ν=2{K}ν=0{K}ν=2{K},\stackrel{{\scriptstyle\nu=-4}}{{\rightarrow}}\{K\downarrow\}\stackrel{{\scriptstyle\nu=-2}}{{\rightarrow}}\{K^{\prime}\downarrow\}\stackrel{{\scriptstyle\nu=0}}{{\rightarrow}}\{K\uparrow\}\stackrel{{\scriptstyle\nu=2}}{{\rightarrow}}\{K^{\prime}\uparrow\}, (72)

and the ν=0\nu=0 gap μZ+(1+12c1214C)V~c\propto\mu_{Z}+(1+{1\over{2}}c_{1}^{2}-{1\over{4}}C)\,\tilde{V}_{c} will open as an enhanced spin gap. Two such filling sequences (71) and (72) appear consistent with part of the (uu-dependent) layer-charge pattern observed in a recent capacitance measurement KFLH .

Let us next consider CR supported by the PZM sector (per spin and valley). Consider a time-dependent uniform electric potential A(t)=A=Ax(t)+iAy(t)A(t)=A=A_{x}(t)+iA_{y}(t), coupled to the PZM sector via the currents R01R^{01} and R1,±2R^{1,\pm 2}, with the Hamiltonian

HA=evζ1AθS𝐩=𝟎01+evζnA(cθS𝟎1n+sθS𝟎0n)+,H_{A}=ev\,\zeta_{1}A_{\theta}\,S_{\bf p=0}^{01}+ev\zeta_{n}A\,(c_{\theta}S_{\bf 0}^{1n}+s_{\theta}S_{\bf 0}^{0n})+\cdots, (73)

where Aθ=cθ2Asθ2AA_{\theta}=c_{\theta}^{2}A-s_{\theta}^{2}A^{{\dagger}}, ζ1=b1rc1κc1\zeta_{1}=b_{1}-r\,c^{\prime}_{1}\approx-\kappa\,c_{1} and ζ±2b±2c1+b±2c1\zeta_{\pm 2}\approx b^{\prime}_{\pm 2}\,c^{\prime}_{1}+b_{\pm 2}\,c_{1}. Accordingly, CR takes place in channels 0θ1θ0_{\theta}\leftarrow 1_{\theta}, 2(1θ,0θ)2\leftarrow(1_{\theta},0_{\theta}), (1θ,0θ)2(1_{\theta},0_{\theta})\leftarrow-2, etc.

Notably, intra PZM resonance is possible BCNM . When orbital mixing is present (u<ucru<u^{\rm cr}), CR arises in the 0θ1θ0_{\theta}\leftarrow 1_{\theta} channel, with excitation energy

ϵexc0θ1θ\displaystyle\epsilon_{\rm exc}^{0_{\theta}\leftarrow 1_{\theta}}\! =\displaystyle= ϵ^0θϵ^1θ(N1N0)𝒜01,\displaystyle\hat{\epsilon}_{0_{\theta}}-\hat{\epsilon}_{1_{\theta}}\!-(N_{1}-N_{0})\,{\cal A}_{01},
=\displaystyle= ϵLs{(1ξ)x+N0q[x]N1q[x]},\displaystyle\!\epsilon_{\rm Ls}\,\big{\{}(1-\xi)x+N_{0}\,q[x]-N_{1}\,q[-x]\big{\}},
q[x]\displaystyle q[x] =\displaystyle= (1/2)(1+x)(13x),\displaystyle(1/2)(1+x)(1-3x), (74)

where xcosθx\equiv\cos\theta, ϵLs=18CV~c\epsilon_{\rm Ls}={1\over{8}}C\,\tilde{V}_{c} and

𝒜01=𝐩v𝐩γ𝐩2G𝐩00G𝐩11=(112c12)V~c12ϵLssin2θ.{\cal A}_{01}\!=\sum_{\bf p}\!v_{\bf p}\gamma_{\bf p}^{2}G^{00}_{\bf-p}G^{11}_{\bf p}=(1-\textstyle{1\over{2}}c_{1}^{2})\,\tilde{V}_{c}-\textstyle{1\over{2}}\epsilon_{\rm Ls}\,\sin^{2}\theta. (75)

On the other hand, when orbital mixing is absent (u>ucr)(u>~{}u^{\rm cr}), the 101\leftarrow 0 channel is activated, with

ϵexc10=ϵ^1ϵ^0+14CV~cN1=ϵLs{ξ1+2N1}.\epsilon_{\rm exc}^{1\leftarrow 0}=\textstyle\hat{\epsilon}_{1}-\hat{\epsilon}_{0}+{1\over{4}}C\tilde{V}_{c}\,N_{1}=\epsilon_{\rm Ls}\,\{\xi-1+2N_{1}\}. (76)

Interestingly, for 0<Nf(=N0)10<N_{\rm f}(=N_{0})\leq 1, there is no correction to CR since |g𝐩10|2|g𝐩00|2+g𝐩11g𝐩00=0|g^{10}_{\bf p}|^{2}-|g^{00}_{\bf p}|^{2}+g^{11}_{\bf p}g^{00}_{\bf-p}=0 holds; see Eq. (27).

Figure 3(b) shows how the excitation spectra evolve with filling NfN_{\rm f} at bias u=0u=0, u=20u=-20 meV<ucr<u^{\rm cr} (thus in valley KK^{\prime}) and u=20u=20 meV >ucr>u^{\rm cr}. While a sizable level gap opens around Nf1N_{\rm f}\sim 1, it is almost cancelled by Coulombic attraction, leaving ϵexc0θ1θ\epsilon_{\rm exc}^{0_{\theta}\leftarrow 1_{\theta}} and ϵexc10\epsilon_{\rm exc}^{1\leftarrow 0} of magnitude of the Lamb shift ϵLsO(0.15V~c)\epsilon_{\rm Ls}\sim O(0.15\,\tilde{V}_{c}) or even smaller.

Such an intra-PZM channel of CR is activated when an orbital band gap develops around ν(±3,±1)\nu\sim(\pm 3,\pm 1). The coupling ζ1κc1\zeta_{1}\approx-\kappa\,c_{1} to the S01S^{01} charge, however, is weak. It will therefore be a challenge to detect CR within the LLL in experiment. On the contrary, CR from or into the LLL serves as a practical probe into the LLL, as discussed in the next section.

VI interband cyclotron resonance

Refer to caption
Figure 4: Cyclotron resonance T2={2PZM,PZM2}T_{2}=\{2\leftarrow{\rm PZM},{\rm PZM}\leftarrow-2\}, with the choice V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 at BB=20 T. (a) Evolution of the level spectra with filling of the PZM sector (Nf=02N_{\rm f}=0\rightarrow 2). Vertical arrows denote the associated CR channels at zero bias u=0u=0. (b) Resonance spectra, in units of ωc\omega_{c} and meV, over the range 0Nf20\lesssim N_{\rm f}\lesssim 2 at zero bias u=0u=0; also shown in scale (01.0)(0\sim 1.0) is the relative strength of each resonance signal. Red circles indicate the most prominent signal at each integer filling. (c) Resonance spectra at bias u=20u=20 meV >ucr(13meV)>u^{\rm cr}(\approx 13{\rm meV}) in valley KK, and at u|K=20u|^{K^{\prime}}=-20 meV <ucr<u^{\rm cr} in valley KK^{\prime}. (d) Evolution of the full T2T_{2} spectra in four {spin,valley} channels with filling of the LLL according to the spin-dominant filling sequence of Eq. (72). (e) Evolution of the T2T_{2} spectra according to the valley-dominant filling sequence of Eq. (71).

VI.1 T2={2T_{2}=\{2\leftarrow PZM, PZM2\leftarrow-2}

Figure 4(a) shows how the level spectra evolve as the PZM sector is gradually filled over 0Nf20~{}\leq N_{\rm f}\leq~{}2 (per spin and valley). The PZM spectra are shifted almost uniformly with bias uu while the n=±2n=\pm 2 spectra are barely affected. As a result, the associated T2T_{2} channels of CR sensitively depend on uu.

For 0u<ucr0\leq u<u^{\rm cr} orbital mixing is present in both valleys and CR arises in four channels, 2(1θ,0θ)2\leftarrow(1_{\theta},0_{\theta}) and (1θ,0θ)2(1_{\theta},0_{\theta})\leftarrow-2 over the range Nf<Nf<Nf+N_{\rm f}^{-}<N_{\rm f}<N_{\rm f}^{+}; see Appendix B for details of the CR spectra. The resonance spectra change in composition and strength in a characteristic way with filling NfN_{\rm f}, as depicted in Fig. 4(b) for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 at u=0u=0 and B=B= 20 T. The strength of each response (function) is proportional to (νiνf)×wmix(\nu_{\rm i}-\nu_{\rm f})\times w_{\rm mix}, i.e, the filling-factor difference between the initial and final levels ×\times mixing weight wmixw_{\rm mix} read from HAH_{A} in Eq. (73). In Fig. 4(b) we also plot such relative weights, {N1cθ2,N0sθ2}\{N_{1}c_{\theta}^{2},N_{0}s_{\theta}^{2}\} for 2{1θ,0θ}2\leftarrow\{1_{\theta},0_{\theta}\} and {(1N1)cθ2,(1N0)sθ2}\{(1-N_{1})c_{\theta}^{2},(1-N_{0})s_{\theta}^{2}\} for {1θ,0θ}2\{1_{\theta},0_{\theta}\}\leftarrow-2. In the spectra red circles indicate the most prominent signal at each integer filling.

The 121\leftarrow-2 and 212\leftarrow 1 channels remain active even for 1<Nf<0-1<N_{\rm f}<0 and 2<Nf<32<N_{\rm f}<3, respectively, i.e., when either the n=2n=-2 or n=2n=2 level is partially filled. The associated CR spectra slightly rise there, because Coulombic attraction diminishes as Nf1N_{\rm f}\rightarrow-1 or Nf3N_{\rm f}\rightarrow 3; the CR signals themselves also vanish in this limit.

When bias uu is increased, e.g., to u=20meV>ucru=20\,{\rm meV}>u^{\rm cr}, orbital mixing disappears in valley KK, with only the 121\leftarrow-2 and 212\leftarrow 1 channels of CR activated, while orbital mixing continues in valley KK^{\prime}, as depicted in Fig. 4(c).

From these spectra one can visualize how the 2spin×2valley=42_{\rm spin}\times 2_{\rm valley}=4 channels (or more) of T2T_{2} compete as the LLL is filled over the range 4ν4-4\leq\nu\leq 4 under a given bias uu. Figure 4(e) illustrates such resonance spectra of u=20u=20 meV at each integer filling, with the valley-dominant filling sequence (uμZ)(u\gg\mu_{\rm Z}) of Eq. (71) assumed. Also Fig. 4(d) shows the T2T_{2} spectra, with the spin-dominant filling sequence (μZ>u+0)\mu_{\rm Z}>u\sim+0) of Eq. (72) assumed. It is clear that, with increasing bias uu, the spectra get split in the valley. Remarkably, the resonance spectra look nearly ee-hh symmetric, i.e., symmetric in ν\nu. An asymmetry is apparent at odd integers ν=(±1,±3)\nu=(\pm 1,\pm 3) in Fig. 4(e), where orbital mixing only arises in valley KK^{\prime}.

It is illuminating here to look into some numbers. From Figs. 4(a)\sim(c) one can read off the following set of level gaps δϵ^m,nϵ^mϵ^n\delta\hat{\epsilon}_{m,n}\equiv\hat{\epsilon}_{m}-\hat{\epsilon}_{n} and excitation energies,

(δϵ^1,2,ϵexc12)|Nf=0\displaystyle(\delta\hat{\epsilon}_{1,-2},\epsilon_{\rm exc}^{1\leftarrow-2})|^{N_{\rm f}=0} u=0\displaystyle\stackrel{{\scriptstyle u=0}}{{\approx}} (105.95,76.92),\displaystyle(105.95,76.92), (77)
(δϵ^2,1,ϵexc21)|Nf=2\displaystyle(\delta\hat{\epsilon}_{2,1},\epsilon_{\rm exc}^{2\leftarrow 1})|^{N_{\rm f}=2} u=0\displaystyle\stackrel{{\scriptstyle u=0}}{{\approx}} (105.08,75.91),\displaystyle(105.08,75.91), (78)
(δϵ^1,2,ϵexc12)|Nf=0\displaystyle(\delta\hat{\epsilon}_{1,-2},\epsilon_{\rm exc}^{1\leftarrow-2})|^{N_{\rm f}=0} u=20\displaystyle\stackrel{{\scriptstyle u=20}}{{\approx}} (93.31,64.51),\displaystyle(93.31,64.51), (79)
(δϵ^2,1,ϵexc21)|Nf=2\displaystyle(\delta\hat{\epsilon}_{2,1},\epsilon_{\rm exc}^{2\leftarrow 1})|^{N_{\rm f}=2} u=20\displaystyle\stackrel{{\scriptstyle u=-20}}{{\approx}} (93.43,64.51)|K,\displaystyle(93.43,64.51)|^{K^{\prime}}, (80)
(δϵ^1,2,ϵexc12)|Nf=0\displaystyle(\delta\hat{\epsilon}_{1,-2},\epsilon_{\rm exc}^{1\leftarrow-2})|^{N_{\rm f}=0} u=20\displaystyle\stackrel{{\scriptstyle u=-20}}{{\approx}} (119.54,90.28)|K,\displaystyle(119.54,90.28)|^{K^{\prime}}, (81)
(δϵ^2,1,ϵexc21)|Nf=2\displaystyle(\delta\hat{\epsilon}_{2,1},\epsilon_{\rm exc}^{2\leftarrow 1})|^{N_{\rm f}=2} u=20\displaystyle\stackrel{{\scriptstyle u=20}}{{\approx}} (117.75,88.94),\displaystyle(117.75,88.94), (82)

in units of meV. These level gaps and CR spectra at NfN_{\rm f}= (0,2) are related via ee-hh conjugation,

(δϵ^1,2,ϵexc12)|Nf=0=(δϵ^2,1,ϵexc21)|μ,d,rNf=2.(\delta\hat{\epsilon}_{1,-2},\epsilon_{\rm exc}^{1\leftarrow-2})|^{N_{\rm f}=0}=(\delta\hat{\epsilon}_{2,1},\,\epsilon_{\rm exc}^{2\leftarrow 1})|_{-\mu,-d,-r}^{N_{\rm f}=2}. (83)

In view of this, Eqs. (77) and (78) imply that the effect of ee-hh breaking (d,r)(d,r) is on the order of 1% in these level and CR spectra. This is further confirmed by the approximate equality between the u=±20u=\pm 20 meV expressions. It is somewhat surprising that the level gaps and CR spectra are only slightly affected by ee-hh breaking (d,r)(d,r) while the level spectra {ϵ^n}\{\hat{\epsilon}_{n}\} themselves are considerably modified (by a few % generally, and much more for ϵ^1\hat{\epsilon}_{1}), as noted in a paragraph below Eq. (60). Accordingly, in experiment, the CR signals of T2T_{2} will look nearly symmetric in ν\nu at even-integer fillings. A shift or splitting of signals with ν\nu around odd-integer filling ν±1\nu\sim\pm 1 and ν±3\nu\sim\pm 3 will reveal the presence or absence of orbital mixing.

It is seen from Eq. (77) that Coulombic attraction accounts for roughly 30% of the relevant level gap here for T2T_{2}. It diminishes rapidly for higher TnT_{n}; it amounts to about 12% for T3T_{3} and 8% for T4T_{4}.

Refer to caption
Figure 5: T3T_{3} spectra (ϵexc32,ϵexc23)(\epsilon_{\rm exc}^{3\leftarrow-2},\epsilon_{\rm exc}^{2\leftarrow-3}), normalized to the zero-thth level gap (ϵ3(0)ϵ2(0))1.18ωc(\epsilon_{3}^{(0)}-\epsilon_{-2}^{(0)})\approx 1.18\,\omega_{c}, for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 at B=B=20 T. (a) Evolution of the T3T_{3} spectra with filling NfN_{\rm f} at bias u=(0,±20)u=(0,\pm 20) meV. Thin dotted lines guide the spectra of the ee-hh symmetric case (d,r)=0(d,r)=0. (b) Evolution of the full T3T_{3} spectra in eight {spin,valley} channels with filling of the LLL according to Eq. (71). (c) For (d,r)=0(d,r)=0 the T3T_{3} spectra are naturally symmetric in ν\nu.

VI.2 Interband resonance T3T5T_{3}\sim T_{5}

Refer to caption
Figure 6: Resonance spectra T4T_{4} and T5T_{5}, normalized to (ϵ4(0)ϵ3(0),ϵ5(0)ϵ4(0)}(1.69,2.13)ωc(\epsilon_{4}^{(0)}\!-\epsilon_{-3}^{(0)},\epsilon_{5}^{(0)}\!-\epsilon_{-4}^{(0)}\}\approx(1.69,2.13)\,\omega_{c}, for V~c/ωc=0.4\tilde{V}_{c}/\omega_{c}=0.4 at BB=20 T. ν\nu refers to the total filling factor, 4(Nf1)4(N_{\rm f}-1), to be realized when NfN_{\rm f} is common to all spins and valleys.

Let us next consider other interband CR, T3T5T_{3}\sim T_{5}. Figure 5(a) shows how the competing T3T_{3} spectra (ϵexc32,ϵexc23)(\epsilon_{\rm exc}^{3\leftarrow-2},\epsilon_{\rm exc}^{2\leftarrow-3}) evolve with filling NfN_{\rm f} of the relevant valley at bias u=(0,±20)u=(0,\pm 20) meV. Thin dotted lines guide the spectra of the (d,r)=0(d,r)=0 case; they are symmetric about Nf=1N_{\rm f}=1 at zero bias u=0u=0. As ee-hh breaking (d,r)(d,r) is turned on, ϵexc32\epsilon_{\rm exc}^{3\leftarrow-2} and ϵexc23\epsilon_{\rm exc}^{2\leftarrow-3} deviate upward and downward, respectively, by roughly 2 % from the symmetric spectra, making the spectra split more on the Nf>1N_{\rm f}>1 side and less on the other.

Figure 5(b) illustrates how the 2×2spin×2valley=82\times 2_{\rm spin}\times 2_{\rm valley}=8 channels of T3T_{3} spectra evolve over the range 4ν4-4\leq\nu\leq 4 at bias u=20u=20 meV, with the valley-dominant filling sequence of Eq. (71) assumed. Splittings among the (ϵexc32,ϵexc23)(\epsilon_{\rm exc}^{3\leftarrow-2},\epsilon_{\rm exc}^{2\leftarrow-3}) spectra, unlike T2T_{2}, develop with ν\nu noticeably, and ee-hh breaking (d,r)(d,r) makes the full spectra visibly asymmetric in ν\nu, in contrast to the ee-hh symmetric [(d,r)=0][(d,r)=0] spectra, depicted in Fig. 5(c).

Figure 6 shows the CR spectra of T4=(ϵexc43,ϵexc34)T_{4}=(\epsilon_{\rm exc}^{4\leftarrow-3},\epsilon_{\rm exc}^{3\leftarrow-4}) and T5=(ϵexc54,ϵexc45)T_{5}=(\epsilon_{\rm exc}^{5\leftarrow-4},\epsilon_{\rm exc}^{4\leftarrow-5}) per spin and valley. Included in the figure are the values of total filling factor ν\nu to be realized when NfN_{\rm f} (except Nf=1N_{\rm f}=1) is common to all valleys and spins, i.e., ν=4(Nf1)\nu=4(N_{\rm f}-1). These higher-energy resonances are practically insensitive to the detailed structure of the LLL and to a bias of u20u\sim 20 meV. Again the CR spectra are less affected by (d,r)0(d,r)\not=0 than level shifts Δϵn\Delta\epsilon_{n}, but are made visibly asymmetric in ν\nu. The competing spectra overlap around Nf1N_{\rm f}\sim-1 or ν8\nu\sim-8 and split more and more as ν6\nu\rightarrow 6 for T4T_{4} and ν12\nu\rightarrow 12 for T5T_{5}.

Presumably, when ν\nu changes over a wide range as in Fig. 6, screening of the Coulomb potential v𝐩v_{\bf p} will become important, as noted KS_CRnewGR ; SL for monolayer graphene. Via screening v𝐩v_{\bf p} will get weaker with increasing |ν||\nu|, making the spectra decrease faster for larger |ν||\nu|.

VII Summary and discussion

Characteristic to bilayer graphene in a magnetic field is an octet of PZM levels nearly degenerate in orbitals n={0,1}n=\{0,1\} as well as in spins and valleys. In this paper we have studied some basic characteristics of such PZM levels and shown that they generally undergo mixing in orbitals {0,1}\{0,1\} as they are gradually filled with electrons. We have examined possible consequences of orbital mixing and how they are detected by an observation of some leading competing channels of interband CR over a finite range of filling factor ν\nu.

It will be illuminating to summarize here why and how orbital mixing arises. Let us first suppose an ee-hh symmetric setting with (d,r)0(d,r)\rightarrow 0 and u0u\rightarrow 0. Coulomb interactions then lead to the orbital Lamb shift. The resulting shift ϵ^0ϵ^1\hat{\epsilon}_{0}-\hat{\epsilon}_{1} is odd under ee-hh conjugation and changes sign as one goes from the empty to filled PZM sector, (ϵ^0ϵ^1)|Nf=0=(ϵ^0ϵ^1)|Nf=2=ϵLs=18CV~c(\hat{\epsilon}_{0}-\hat{\epsilon}_{1})|^{N_{\rm f}=0}=-(\hat{\epsilon}_{0}-\hat{\epsilon}_{1})|^{N_{\rm f}=2}=\epsilon_{\rm Ls}={1\over{8}}C\tilde{V}_{c}; this is because the zero-modes are ee-hh selfconjugate [so that ϵ^n|Nf=0=ϵ^n|XNf=2\hat{\epsilon}_{n}|^{N_{\rm f}=0}=-\hat{\epsilon}_{n}|_{-X}^{N_{\rm f}=2} for n=(0,1)n=(0,1)]. It is this level inversion that drives orbital mixing with filling of the PZM sector, as discussed in Sec. V. When valley and ee-hh breaking (u,d,r)(u,d,r) is turned on, the level shift takes a modified form (ϵ^0ϵ^1)|Nf=(±1ξ)ϵLs(\hat{\epsilon}_{0}-\hat{\epsilon}_{1})|^{N_{\rm f}}=(\pm 1-\xi)\epsilon_{\rm Ls} at Nf=(0,2)N_{\rm f}=(0,2), with ξ=O(μ,d,r)\xi=O(\mu,d,r), and begins to change with bias uu slightly and almost linearly. The bias uu thereby acquires a critical value ucru^{\rm cr}, beyond which orbital mixing disappears. In this way, orbital mixing is driven by the orbital Lamb shift, and generally takes place in either valley or both (per spin) as bias uu is varied.

In our analysis, special attention has been paid to ee-hh conjugation and valley-interchange operations, which govern, as in Eqs. (39) and (43), the level and CR spectra in bilayer graphene. In experiment, Coulombic corrections will be seen as variations of the interband CR spectra with filling ν\nu, ee-hh breaking as an asymmetry of the spectra about ν=0\nu=0, and valley breaking as variations or splitting of the spectra with bias uu. Orbital mixing will be detected by an opening of some additional channels of CR. The T2T_{2} channels of interband CR, in particular, will serve as a direct and most sensitive probe to explore the novel characteristics of the LLL in bilayer graphene.

Appendix A Counterterms δϵn\delta\epsilon_{n}

In Sec. IV the bare level spectra are written as ϵn=ϵnren+δϵn\epsilon_{n}=\epsilon_{n}^{\rm ren}+\delta\epsilon_{n}. In this appendix we outline how to calculate the counterterms δϵn\delta\epsilon_{n} numerically. For given ϵnren\epsilon_{n}^{\rm ren}, one can write the associated counterterms as

δϵn=δctϵnrenλnδv/vren,\delta\epsilon_{n}=\delta_{\rm ct}\epsilon_{n}^{\rm ren}\equiv\lambda_{n}\,\delta v/v^{\rm ren}, (84)

with the differential operator

δct=δvvren+δγ1γ1ren+δΔΔren\delta_{\rm ct}=\delta v{\partial\over{\partial v^{\rm ren}}}+\delta\gamma_{1}{\partial\over{\partial\gamma_{1}^{\rm ren}}}+\delta\Delta{\partial\over{\partial\Delta^{\rm ren}}} (85)

acting on ϵnren\epsilon_{n}^{\rm ren} and with (δv,δγ1,δΔ)(\delta v,\delta\gamma_{1},\delta\Delta) defined in Eq. (46). This formula allows one to evaluate λn\lambda_{n} analytically. Alternatively, one can let δct\delta_{\rm ct} act on the reduced matrix ^N\hat{\cal H}_{N} in Eq. (11), and write

δϵn=(𝐛n)δct^N𝐛n=λnδv/vren.\delta\epsilon_{n}=({\bf b}_{n})^{{\dagger}}\!\cdot\delta_{\rm ct}\hat{\cal H}_{N}\!\cdot{\bf b}_{n}=\lambda_{n}\,\delta v/v^{\rm ren}. (86)

Actually, ΛNδct^N/(δv/vren)\Lambda_{N}\equiv\delta_{\rm ct}\hat{\cal H}_{N}/(\delta v/v^{\rm ren}) is given by ^N\hat{\cal H}_{N} with substitution (μ,r)0(\mu,r)\rightarrow 0 first and subsequently g(g+rd)h(r)g\rightarrow(g+r\,d)\,h(r) and d2(d+rg)h(r)d\rightarrow 2(d+r\,g)h(r). In this way one can directly calculate λn=𝐛nΛN𝐛n\lambda_{n}={\bf b}_{n}^{{\dagger}}\!\cdot\!\Lambda_{N}\!\cdot\!{\bf b}_{n} from eigenmodes 𝐛n{\bf b}_{n}.

Appendix B Resonance Spectra T2T_{2}

In this Appendix, we outline the derivation of the CR spectra T2={2PZM,PZM2}T_{2}=\{2\leftarrow{\rm PZM},\ {\rm PZM}\leftarrow-2\} examined in Sec. VI. Let us first write the renormalized level spectra as ϵ^n=ϵ^n|Nf=0+δϵnSE\hat{\epsilon}_{n}=\hat{\epsilon}_{n}|^{N_{\rm f}=0}+\delta\epsilon_{n}^{\rm SE}. As the PZM sector is gradually filled over the range 0Nf=N1+N020\leq N_{\rm f}=N_{1}+N_{0}\leq 2, each level nn acquires an additional self-energy correction

δϵnSE=𝐩v𝐩γ𝐩2{N1|G𝐩n1|2+N0|G𝐩n0|2}.\delta\epsilon_{n}^{\rm SE}=-\sum_{\bf p}v_{\bf p}\gamma_{\bf p}^{2}\,\{N_{1}|G^{n1}_{\bf p}|^{2}+N_{0}|G^{n0}_{\bf p}|^{2}\}. (87)

Direct calculation of δϵ0SE\delta\epsilon_{0}^{\rm SE} and δϵ1SE\delta\epsilon_{1}^{\rm SE} leads to (ϵ^0θ,ϵ^1θ)(\hat{\epsilon}_{0_{\theta}},\hat{\epsilon}_{1_{\theta}}) in Eq. (62). Adding Coulombic attraction terms

𝒜nj=𝐩v𝐩γ𝐩2G𝐩nnG𝐩jj{\cal A}_{nj}=\sum_{\bf p}v_{\bf p}\gamma_{\bf p}^{2}\,G^{nn}_{\bf p}G^{jj}_{\bf-p} (88)

then yields the CR spectra ϵexcnj=ϵ^nϵ^j(νjνn)𝒜n,j\epsilon_{\rm exc}^{n\leftarrow j}=\hat{\epsilon}_{n}-\hat{\epsilon}_{j}-(\nu_{j}-\nu_{n})\,{\cal A}_{n,j}.

When orbital mixing is present, i.e., for u<ucru<u^{\rm cr}, interband CR, T2T_{2}, hosts four active channels (per spin and valley) over the range Nf<Nf<Nf+N_{\rm f}^{-}<N_{\rm f}<N_{\rm f}^{+}. Associated with the n=1θn=1_{\theta} level are the excitation spectra,

ϵexc21θ\displaystyle\epsilon_{\rm exc}^{2\leftarrow 1_{\theta}} =\displaystyle= ϵ^2ϵ^1θN1𝒜21,\displaystyle\hat{\epsilon}_{2}-\hat{\epsilon}_{1_{\theta}}-N_{1}\,{\cal A}_{21}, (89)
ϵexc1θ2\displaystyle\epsilon_{\rm exc}^{1_{\theta}\leftarrow-2} =\displaystyle= ϵ^1θϵ^2(1N1)𝒜1,2,\displaystyle\hat{\epsilon}_{1_{\theta}}-\hat{\epsilon}_{-2}\!-(1-N_{1})\,{\cal A}_{1,-2}, (90)

and those associated with the n=0θn=0_{\theta} level are

ϵexc20θ\displaystyle\epsilon_{\rm exc}^{2\leftarrow 0_{\theta}} =\displaystyle= ϵ^2ϵ^0θN0𝒜20,\displaystyle\hat{\epsilon}_{2}-\hat{\epsilon}_{0_{\theta}}-N_{0}\,{\cal A}_{20}, (91)
ϵexc0θ2\displaystyle\epsilon_{\rm exc}^{0_{\theta}\leftarrow-2} =\displaystyle= ϵ^0θϵ^2(1N0)𝒜0,2.\displaystyle\hat{\epsilon}_{0_{\theta}}-\hat{\epsilon}_{-2}-(1-N_{0})\,{\cal A}_{0,-2}. (92)

In actual calculations one can simplify, for n(0,1)n\not=(0,1), |Gn1|2cθ2|gn1|2+sθ2|gn0|2|G^{n1}|^{2}\rightarrow c_{\theta}^{2}|g^{n1}|^{2}+s_{\theta}^{2}|g^{n0}|^{2}, GnnG11gnn(cθ2g11+sθ2g00)G^{nn}G^{11}\rightarrow g^{nn}(c_{\theta}^{2}\,g^{11}+s_{\theta}^{2}\,g^{00}), etc., under symmetric integration 𝐩\sum_{\bf p}. Evaluating Eqs. (89) \sim (92) numerically leads to the excitation spectra depicted in Fig. 4.

References

  • (1) E. McCann and V. I. Fal’ko, Phys. Rev. Lett. 96, 086805 (2006).
  • (2) K. S. Novoselov, E. McCann, S. V. Morozov, V. I. Fal’ko, M. I. Katsnelson, U. Zeitler, D. Jiang, F. Schedin, and A. K. Geim, Nat. Phys. 2, 177 (2006).
  • (3) T. Ohta, A. Bostwick, T. Seyller, K. Horn, and E. Rotenberg, Science 313, 951 (2006).
  • (4) A. J. Niemi and G. W. Semenoff, Phys. Rev. Lett. 51, 2077 (1983).
  • (5) Y. Barlas, R. Côté, K. Nomura, and A. H.  MacDonald, Phys. Rev. Lett. 101, 097601 (2008).
  • (6) K. Shizuya, Phys. Rev. B 79, 165402 (2009).
  • (7) Y. Barlas, R. Côté, J. Lambert, and A. H.  MacDonald, Phys. Rev. Lett. 104, 096802 (2010).
  • (8) R. Côté, J. Lambert, Y. Barlas, and A. H. MacDonald, Phys. Rev. B 82, 035445 (2010).
  • (9) R. Côté, W. Luo, B. Petrov, Y. Barlas, and A. H. MacDonald, Phys. Rev. B 82, 245307 (2010).
  • (10) R. Nandkishore and L. Levitov, Phys. Rev. B 82, 115124 (2010).
  • (11) M. Mucha-Kruczyn´\acute{\rm n}ski, I. L. Aleiner, and V. I. Fal’ko, Phys. Rev. B 84, 041404(R) (2011).
  • (12) K. Shizuya, Phys. Rev. B 86, 045431 (2012).
  • (13) B. E. Feldman, J. Martin and A. Yacoby, Nature Phys. 5, 889 (2009).
  • (14) Y. Zhao, P. Cadden-Zimansky, Z. Jiang, and P. Kim, Phys. Rev. Lett. 104, 066801 (2010).
  • (15) R. T. Weitz, M. T. Allen, B. E. Feldman, J. Martin and A. Yacoby, Science 330, 812 (2010).
  • (16) A. S. Mayorov, D. C. Elias, M.Mucha-Kruczynski, R. V. Gorbachev, T. Tudorovskiy, A. Zhukov, S. V. Morozov, M. I. Katsnelson, V. I. Fal’ko, A. K. Geim, K. S. Novoselov, Science 333, 860 (2011).
  • (17) J. Velasco Jr., et al., Nat. Nanotechnol. 7, 156 (2012).
  • (18) J. Martin, B. E. Feldman, R. T. Weitz, M. T. Allen, and A. Yacoby, Phys. Rev. Lett. 105, 256806 (2010).
  • (19) A. Kou, B. E. Feldman, A. J. Levin, B. I. Halperin, K. Watanabe, T. Taniguchi, and A. Yacoby, Science 345, 55 (2014).
  • (20) B. M. Hunt, J. I. A. Li, A. A. Zibrov, L. Wang, T. Taniguchi, K. Watanabe, J. Hone, C. R. Dean, M. Zaletel, R. C. Ashoori, and A. F. Young, Nat. Commun. 8, 948 (2017).
  • (21) D. S. L. Abergel and V. I. Fal’ko, Phys. Rev. B 75, 155430 (2007).
  • (22) Z. Jiang, E. A. Henriksen, L. C. Tung, Y.-J. Wang, M. E. Schwartz, M. Y. Han, P. Kim, and H. L. Stormer, Phys. Rev. Lett. 98, 197403 (2007).
  • (23) R. S. Deacon, K.-C. Chuang, R. J. Nicholas, K. S. Novoselov, and A. K. Geim, Phys. Rev. B 76, 081406(R) (2007).
  • (24) E. A. Henriksen, P. Cadden-Zimansky, Z. Jiang, Z. Q. Li, L.-C. Tung, M. E. Schwartz, M. Takita, Y.-J. Wang, P. Kim, and H. L. Stormer, Phys. Rev. Lett. 104, 067404 (2010).
  • (25) E. A. Henriksen, Z. Jiang, L.-C. Tung, M. E. Schwartz, M. Takita, Y.-J.Wang, P. Kim, and H. L. Stormer, Phys. Rev. Lett. 100, 087403 (2008).
  • (26) M. Orlita, C. Faugeras, J. Borysiuk, J. M. Baranowski, W. Strupiński, M. Sprinkle, C. Berger, W. A. de Heer, D. M. Basko, G. Martinez, and M. Potemski, Phys. Rev. B 83, 125302 (2011).
  • (27) A. Iyengar, J. Wang, H. A. Fertig, and L. Brey, Phys. Rev. B 75, 125430 (2007).
  • (28) Yu. A. Bychkov, and G. Martinez, Phys. Rev. B 77, 125417 (2008).
  • (29) K. Shizuya, Phys. Rev. B 81, 075407 (2010).
  • (30) R. Roldán, J.-N. Fuchs, and M. O. Goerbig, Phys. Rev. B 82, 205418 (2010).
  • (31) K. Shizuya, Phys. Rev. B 84, 075409 (2011).
  • (32) J. Sári and C. Töke, Phys. Rev. B 87, 085432 (2013).
  • (33) K. Shizuya, Phys. Rev. B 98, 115419 (2018).
  • (34) A. A. Sokolik and Y. E. Lozovik, Phys. Rev. B 99, 085423 (2019).
  • (35) B. J. Russell, B. Zhou, T. Taniguchi, K. Watanabe, and E. A. Henriksen, Phys. Rev. Lett. 120, 047401 (2018).
  • (36) J. Pack, B. J. Russell, Y. Kapoor, J. Balgley, J. Ahlers, T. Taniguchi, K. Watanabe, and E. A. Henriksen, preprint arXiv:2001.08879 [cond-mat.mes-hall].
  • (37) B. Hunt, J. D. Sanchez-Yamagishi, A. F. Young, M. Yankowitz, B. J. Leroy, K. Watanabe, T. Taniguchi, P. Moon, M. Koshino, P. Jarillo-Herrero, and R. C. Ashoori, Science 340, 1427 (2013).
  • (38) L. M. Zhang, Z. Q. Li, D. N. Basov, and M. M. Fogler, Z. Hao, and M. C. Martin, Phys. Rev. B 78, 235408 (2008).
  • (39) Z. Q. Li, E. A. Henriksen, Z. Jiang, Z. Hao, M. C. Martin, P. Kim, H. L. Stormer, andD. N. Basov, Phys. Rev. Lett. 102, 037403 (2009).
  • (40) J. Nilsson, A. H. Castro Neto, F. Guinea, and N. M. R. Peres, Phys. Rev. B 78, 045405 (2008).
  • (41) J. Jung and A. H. MacDonald, Phys. Rev. B 89, 035405 (2014).
  • (42) In ref. KS_CR_eh we adopted a slightly different set of band parameters. All our numerical analysis in the present paper has also been made with it and essentially the same conclusion has been reached.
  • (43) S. M. Girvin, A. H. MacDonald, and P. M. Platzman, Phys. Rev. B 33, 2481 (1986).
  • (44) K. Shizuya, Int. J. Mod. Phys. B 33, 1950171 (2019).
  • (45) C. Kallin and B. I. Halperin, Phys. Rev. B 30, 5655 (1984).
  • (46) A. H. MacDonald, H. C. A. Oji, and S. M. Girvin, Phys. Rev. Lett. 55, 2208 (1985).